content
stringlengths 1
15.9M
|
---|
\section{Introduction}
Substantial progress has been made towards the implementation of
coherent interfaces between an electron
spin in a quantum dot and photons \cite{mikkelsen_science2008, berezovsky_science2008, yilmaz_prl2010, Liu_adv2010}.
The coupling of single spins to photons is a promising mechanism for implementing
quantum information processing schemes based on quantum dots, where
a qubit (quantum bit of information) is represented by electron spin degrees of freedom
\cite{loss_divicenzo_proposaL_QD}.
The main source of electron spin decoherence in semiconductor quantum dots is considered
to be the hyperfine interaction to the nuclear spins of the lattice atoms
\cite{khaetskii_prl2002,coish_statsolidi2009}.
The dephasing time due to the interaction with an unprepared ensemble of nuclear spins is about 10 ns
\cite{petta_science2005, coish_statsolidi2009}, which is much shorter than the single qubit operation
times by means of electronic control ($\approx$ 100 ns \cite{koppens_nature2006}).
There are some ways to deal with this problem: to prolong the decoherence time
using advanced spin-echo techniques \cite{bluhm_natphys2010} or dynamical nuclear polarization
\cite{foletti_natphys2009}, to use nuclear spin ``poor" materials such as
graphene \cite{trauzettel_natphys2007} or nitrogen-vacancy centers in diamond \cite{hanson_nature2008}
or to shorten the qubit manipulation time \cite{hugo_2010}. The optical driving of
electron spins has this last advantage and offers mechanisms for single qubit operation time around
30 ps \cite{berezovsky_science2008}. Besides single qubit manipulations, two qubit operations are
required for implementing quantum algorithms.
One of the schemes for optical exchange interaction between
two spins is based on virtual Raman transitions between valence band heavy holes and
conduction band electrons \cite{imamoglu_prl99}.
Thereby, a set of quantum dots is coupled to a high finesse cavity.
Each quantum dot is doped with a single electron and
can be controlled by linearly polarized laser light.
The laser polarization is perpendicular
to the growth direction of the quantum dots and
perpendicular to the cavity field polarization as well,
so that laser light is acting on the quantum dots without
driving the cavity. A magnetic field is applied in the Voigt configuration, i.e.,
perpendicular to the growth direction and parallel to the cavity field polarization,
so that single photons emited by quantum dots escape
into the cavity and can be detected afterwards.
The cavity frequency is chosen such that it is
slightly detuned from the transition from
the electron spin down state to the trion state,
while the laser frequency is detuned from the
transition from the spin up state to the trion state.
If these detunings are the same then the $\Lambda$ system is
in a two-photon resonance that allows for manipulating a single electron spin
(Figs.~\ref{schematic} and \ref{scheme}).
If the detunings in a quantum dot are different, but the difference of the detunings
(the detuning from the two-photon resonance) match
for a pair of quantum dots, then the electron spins in these quantum dots can
resonantly exchange energy through virtual cavity photons (Fig. \ref{schematic}).
In this way, virtual photons can be used
for performing coherent interaction between arbitrary qubits.
Two-qubit operations
can potentially be carried out on a subnanosecond timescale \cite{imamoglu_prl99},
therefore the decoherence can be expected to originate mainly from
nuclear spins.
In this paper, we calculate the fidelities of such two-qubit operations
in the presence of the hyperfine interaction.
We will study in particular partial SWAP operations and the controlled-not (CNOT) gate.
For this purpose, we first calculate the fidelity of the quantum gates of interest
analytically to second order of the Overhauser field and subsequently we perform
an exact numerical calculation for comparison.
\begin{figure}
\includegraphics[width=0.45\textwidth]{cavity}
\caption{Cavity mode mediated spin-spin interaction. If the difference of the
detunings $\Delta$ for two quantum dots coincides, then they can exchange virtual cavity photons of
energy $\omega_c$. Here, $\omega_c$ and $\omega_L^{i}$ are the frequencies of the cavity mode
and the laser acting on quantum dot $i$, and $\Delta$ denotes the difference between the detunings,
i.e. the detuning from the two-photon resonance on a given quantum dot.}
\label{schematic}
\end{figure}
\section{Model}
We start from the microscopic model describing a quantum dot embedded in a cavity and
irradiated by a laser field,
\begin{eqnarray}
H &=& \sum_{\sigma =\uparrow,\downarrow,\pm 3/2}\omega_\sigma e_\sigma^\dagger e_\sigma
+\omega_c a^\dagger_c a_c + \omega_L a^\dagger_L a_L + H_{\rm hf}\nonumber\\
& & + g_c a_c^\dagger e_{v}^\dagger e_{\uparrow} - ig_L a_L^\dagger
e_{v}^\dagger e_{\downarrow} + h.c.,
\label{eh-Hamiltonian}
\end{eqnarray}
where $\omega_{\sigma}$ and $e_\sigma$ are the energy and annihilation operators for
the electronic states in the conduction band with spin $\sigma = \pm 1/2 = \uparrow, \downarrow$
and in the valence band with total angular momentum $\sigma = \pm 3/2$ (we assume that the
heavy hole subband is sufficiently split from the light hole band to allow for pure heavy hole
excitation).
The quantization axis for the electrons is set by the external magnetic field along the
$x$ direction, while the quantization axis in the valence band is given by the structure and
assumed to be in $z$ direction. Furthermore, for a linearly
polarized cavity mode ($\omega_c$, $a_c$) along the $x$ direction and
linearly polarized laser mode ($\omega_L$,$a_L$) in the $y$ direction, we obtain
a radiative coupling to the linear combination $e_v = (e_{-3/2}-e_{3/2})/\sqrt{2}$ in the
valence band. The coupling strengths to the cavity and laser modes
are denoted with $g_c$ and $g_L$, respectively.
The hyperfine interaction of a conduction band electron in the quantum
dot with the surrounding nuclear spins can be written as
\begin{equation}
H_{\rm hf} = {\bf S}\cdot \sum_{k=1}^N A_k {\bf I}^k,
\end{equation}
where ${\bf I}^k$ is the operator belonging to the $k$-th nuclear
spin in contact with the electron and $A_k$ denotes the corresponding
hyperfine coupling constant.
The average hyperfine coupling constant for GaAs is 90 $\mu$eV
\cite{cerletti_nano2005}.
The electron spin operator is given by
${\bf S} = \tfrac12\sum_{\sigma,\sigma'=\uparrow,\downarrow} e_\sigma^\dagger
\boldsymbol{\sigma}_{\sigma\sigma'}e_{\sigma'}$, where
$\boldsymbol{\sigma}=(\sigma_x,\sigma_y,\sigma_z)$
is the vector consisting of Pauli matrices.
Here, we neglect the dipolar hyperfine coupling of the valence band
states since it is typically smaller.
In the presence of a magnetic field along the $x$ direction exceeding
the nuclear field of typically $\sim 10\,{\rm mT}$, we can neglect the
flip-flop terms and write
$H_{\rm hf}\approx S_x \sum_{k=1}^N A_k I_x^k \equiv S_x h$,
where $h$ denotes the Overhauser (nuclear) field operator in $x$
direction. We do not take the nuclear Zeeman terms into account, because
they are considerably smaller than the thermal energy and the electron Zeeman
energy.
We can then combine the hyperfine Hamiltonian with the first term in
Eq.~(\ref{eh-Hamiltonian}) by using
$\omega_\sigma = (g_e \mu_B B + h ) \sigma /2$
for $\sigma=\uparrow,\downarrow=\pm 1$,
where $B$ is the magnetic field applied along $x$
and $g_e$ is the effective electron g-factor.
We can now replace the operator $h$ with one of its eigenvalues
and perform an average over $h$ later. This allows us to
follow the steps performed in Ref.~\cite{imamoglu_prl99}
before taking the average over nuclear configurations.
In GaAs quantum dots the number $N$ of nuclear spins is
large, typically between $10^5$ and $10^6$, and therefore
the Overhauser field follows a Gaussian distribution
around mean value 0 and with variance $\sigma\simeq A/\sqrt{N}$ \cite{cerletti_nano2005}.
In order to eliminate the valence band states, we perform a
Schrieffer-Wolff transformation
$H_{\rm eff} = e^{-S}He^S$ with the anti-Hermitian operator
\begin{equation}
S = \frac{g_c}{\omega_\uparrow-\omega_v-\omega_c} a^\dagger_c e^\dagger_ve_\uparrow
- i\frac{g_L}{\omega_\downarrow-\omega_v-\omega_L} a^\dagger_L
e^\dagger_v e_\downarrow
- h.c.,
\end{equation}
where we use $\omega_{3/2}\approx\omega_{-3/2}\equiv \omega_v$.
By setting the valence band occupation number to $e^\dagger_v e_v\approx 1$ and
using that the laser field is in a coherent state with $g_L a_L\approx \Omega_L
\exp(-i\omega_L t)$, we obtain to second order in the cavity and
laser couplings $g_c$ and $g_L$,
\begin{eqnarray}
H_{\rm eff} = \omega_c a_c^\dagger a_c &+& \sum_{i=1,2} \Bigg[
(g_e\mu_B B + h_i)\sigma^i_{\uparrow\uparrow} \label{Heff}\\
& & +i g_{\rm eff}^i (a^\dagger_c \sigma^i_{\downarrow\uparrow}e^{-i\omega_{L}^i t} - h.c.)\nonumber\\
& &
-\frac{g_c^2}{\Delta_c^i(h_i)}\sigma_{\downarrow\downarrow}^i
a_c^\dagger a_c
-\frac{(\Omega_L^i)^2}{\Delta_L^i(h_i)}\sigma_{\uparrow\uparrow}^i\Bigg],\nonumber
\end{eqnarray}
where the index $i$ refers to the respective quantities in
dot $i=1,2$, and $\sigma_{i\,j}=| i\rangle\langle
j|$, $i, j=\uparrow,\downarrow$.
The first term inside the sum in Eq.~(\ref{Heff}) is the Zeeman
splitting due to the external and nuclear fields, the second term describes
the effective coupling of the cavity mode to the quantum dot electron
spins with strength
\begin{equation}
g_{\rm eff}^i(t)=\frac{g_c\Omega^i_L(t)}{2}\left(\frac{1}{\Delta^i_c(h_i)}+\frac{1}{\Delta^i_L(h_i)}\right),
\label{geff}
\end{equation}
where
\begin{eqnarray}
\Delta_c^i(h_i) &=& \frac{g_e\mu_B B + h_i}{2}-\omega_v-\omega_c =
\Delta_c^i +\frac{h_i}{2},\label{Deltac}\\
\Delta_L^i (h_i)&=& -\frac{g_e\mu_B B + h_i}{2}-\omega_v-\omega_L^i =
\Delta_L^i -\frac{h_i}{2} ,\label{DeltaL}
\end{eqnarray}
are the detunings of the cavity and laser fields (s. Fig. \ref{scheme}),
and the last two terms in Eq. (\ref{Heff}) can be interpreted as the Lamb and Stark shifts
of the cavity and quantum dot levels, respectively.
\begin{figure}
\includegraphics[width=0.35\textwidth]{scheme}
\caption{Energy level scheme for a quantum dot filled with a single electron
and coupled to a cavity mode.
The Zeeman-split single-electron states can be excited to a trion (negatively charged exciton) state
$|X^{-}\rangle$
by coupling to the cavity or laser field.
Both cavity and laser field frequencies are detuned by $\Delta_c$ and $\Delta_L$
from resonance, and the combined system
is detuned from its two-photon resonance by $\Delta = \Delta_c-\Delta_L$.
The Overhauser shift caused by the hyperfine coupling to the nuclear spins leads to
a fluctuating detuning, thus reducing the fidelity of the optically generated quantum gates.}
\label{scheme}
\end{figure}
A second Schrieffer-Wolff transform can be used to also eliminate the
cavity mode, which leads to
the effective photon-mediated interaction between two spins $i$ and $j$
in the interaction picture with \\
$H_0=\sum_i (g_e\mu_B B+h_i)\sigma^i_{\uparrow\uparrow}$ ,
\begin{equation}
H_{\rm int}^{ij} = \frac{\tilde{g}_{ij}(t)}{2}(\sigma^i_{\uparrow\downarrow} \sigma^j_{\downarrow\uparrow}{\rm e}^{i(h_i-h_j)t}
+ \sigma^j_{\uparrow\downarrow} \sigma^i_{\downarrow\uparrow}{\rm e}^{-i(h_i-h_j)t}),
\label{Hint}
\end{equation}
where
\begin{equation}
\tilde{g}_{ij}(t)=\frac{g^i_{\rm eff}(t)g^j_{\rm eff}(t)}{2}\left(\frac{1}{\Delta_i(h_i)}+\frac{1}{\Delta_j(h_j)}\right),
\label{gij}
\end{equation}
with
\begin{equation}
\Delta_i (h_i)= \Delta^i_c (h_i)-\Delta^i_L (h_i) = \Delta_i + h_i,
\label{Delta}
\end{equation}
represents the coupling strength between
two electron spins.
The time-independent interaction Hamiltonian Eq.~(\ref{Hint}) strictly
applies to the two-photon resonance $\Delta_i = \Delta_j \equiv \Delta$ in
the absence of nuclear spins. By going over into
the rotating frame with nuclear spins by
$R=\exp{(i\,t(h_i\sigma^i_{\uparrow\uparrow}+h_j\sigma^j_{\uparrow\uparrow}))}$
we obtain a time-independent effective interaction Hamiltonian with Heisenberg transverse coupling type between two electron spins:
\begin{equation}
\tilde{H}^{\rm int}_{ij}=R^{\dagger}H_{\rm int}^{ij} R= \frac{\tilde{g}_{ij}(t)}{2}(\sigma^i_y\sigma^j_y+\sigma^i_z\sigma^j_z).
\end{equation}
The unitary time evolution operator of the interaction between two electron spins is
\begin{equation}
\tilde{U}(\phi) = \exp\left[-i \int \tilde{H}_{\rm int}^{ij} dt\right],
\label{U}
\end{equation}
with $\phi = \int \tilde{g}_{ij}(t)\, dt$.
The time evolution operator in original frame is then
\begin{eqnarray}
U(\phi)&=&R \,\tilde{U}(\phi)R^{\dagger}=\hspace{2.5cm} \\
&=&\begin{pmatrix}
1 & 0& 0& 0\\
0 & \cos (\phi) & i\sin(\phi) {\rm e}^{i(h_i-h_j)t} & 0\\
0 & i\sin(\phi) {\rm e}^{-i(h_i-h_j)t}& \cos(\phi)& 0\\
0 & 0 & 0 & 1
\end{pmatrix}. \nonumber
\end{eqnarray}
In the subspace $\{|\uparrow\downarrow\rangle,\,|\downarrow\uparrow\rangle\} $ the operator $U(\phi)$ acts as
a rotation $\exp(i\,\phi\,\boldsymbol{\sigma}\cdot\hat{n})$ with $\hat{n}=(-\cos\{(h_i-h_j)t\}, \sin\{(h_i-h_j)t\}, 0)$.
Thus the changed time evolution of the two-electron spin state due to hyperfine interaction can be interpreted as a modified rotation in the mentioned subspace (s. Fig. \ref{SP}).
There are two distinct effects due to the nuclear spins. First, the interaction phase becomes Overhauser field dependent $\phi=\phi(h_i, h_j)$ and, secondly, the rotation axis starts to precess
in the x-y plane with $(h_i-h_j)t$. The second effect has the maximal contribution when $\phi(0)=m\tfrac\pi2$, where
$\phi(0)=\tilde{g}_{ij}(h_i=0, h_j=0)\, t\equiv \tilde{g}_{ij}(0)\, t$ and where $m$ is an integer. When $\phi(0)=m\pi$, the trajectory of the two-electron spin state affected by nuclear spins coincides with an unaffected one.
\begin{figure}
\includegraphics[width=0.35\textwidth]{sphere}
\caption{Effect of the Overhauser fields $h_i$ and $h_j$ on the time evolution of the two-electron spin state:
change of the interaction phase and precessing of the rotation axis in the x-y plane with $(h_i-h_j)t$}.
\label{SP}
\end{figure}
The two-qubit CNOT operation can be implemented as a sequence of single spin rotations combined
with the unitary time evolution operator \cite{imamoglu_prl99}:
\begin{equation}
\begin{split}
U_{\rm CNOT}=\mathrm{e}^{-i\tfrac\pi4 \sigma^j_z}\, \mathrm{e}^{-i \tfrac\pi4}\,
\mathrm{e}^{-i\tfrac\pi3\hat{n}_i\cdot\boldsymbol{\sigma}_i} \,
\mathrm{e}^{-i\tfrac\pi3\hat{n}_j\cdot\boldsymbol{\sigma}_j}\times \\
\times U\left(\tfrac\pi4\right)\, \mathrm{e}^{-i\tfrac\pi2\sigma^i_z} \, U\left(\tfrac\pi4\right)\,
\mathrm{e}^{-i\tfrac\pi4\sigma^i_y}\, \mathrm{e}^{-i\tfrac\pi4\sigma^j_y} \, \mathrm{e}^{i\tfrac\pi4\sigma^j_z},
\end{split}
\end{equation}
where $\hat{n}_i=(1, 1, -1)/\sqrt{3}$, $\hat{n}_j=(-1, 1, 1)/\sqrt{3}$.
We calculate the fidelity of the generated unitary gate
in the presence of the Overhauser field with respect to the ideal unitary gate
(without Overhauser field) in order to quantify the effect of the nuclear spins.
This fidelity reflects the difference in final state after the gate operation, averaged
over pure input quantum states \cite{petersen_pra2007}.
The average
fidelity for operators acting in 4-dimensional Hilbert space is given by \cite{ghoish_pra2010}
\begin{equation}
F(\hat{O}(0), \hat{O}(h))=\frac{4+|\mbox{Tr}(\hat{O}^{\dagger}(0)\hat{O}(h))|^2}{20},
\label{generalF}
\end{equation}
where $\hat{O}(h)$ and $\hat{O}(0)$ denote the unitary operator with and without Overhauser field dependence.
Next, the average fidelity Eq.~(\ref{generalF}) must be averaged over the nuclear spin field distributions in
the quantum dots $i$ and $j$ to extract the Overhauser field dependence of the fidelities for the time
evolution and the CNOT operator.
\section{Fidelities in second order hyperfine interaction}
Starting from the microscopic model Eq.~(\ref{eh-Hamiltonian}) and
performing two subsequent Schrieffer-Wolff transformations
\cite{imamoglu_prl99}
while including the hyperfine interaction, we find the effective
spin-spin interaction Hamiltonian as described above.
Using Eqs.~(\ref{geff}), (\ref{Deltac}), (\ref{DeltaL}),
(\ref{gij}), and (\ref{Delta}), we find that the effective spin-spin
coupling depends on the Overhauser fields $h_i$ and $h_j$ in the quantum dots,
\begin{eqnarray}
\label{gfull}
\tilde{g}_{ij}(h_i,h_j)&=&\frac{g^2_{c}\Omega^i_L\Omega^j_L}{16}
\left(\frac{1}{\Delta_i+h_i}+\frac{1}{\Delta_j+h_j}\right)\nonumber\\
&&\times\left(\frac{1}{\Delta_c^i+h_i/2}+
\frac{1}{\Delta^i_L-h_i/2}\right)\\
&&\times\left(\frac{1}{\Delta_c^j+h_j/2}+\frac{1}{\Delta^j_L-h_j/2}\right).\nonumber
\end{eqnarray}
We assume that the fluctuations of the Overhauser field around its zero mean value are small,
since $h\sim A\sim 10^{-5}$ eV. Therefore we can investigate analytically the fidelities to the second
order of the Overhauser field. Since the interaction Hamiltonian (\ref{Hint}) is nuclear field
dependent partly through the coupling coefficient
$\tilde{g}_{ij}$, we need to expand it with respect to the Overhauser field in both coupled quantum dots,
\begin{eqnarray}
\frac{\tilde{g}_{ij}(h_i,h_j)}{\tilde{g}_{ij}(0)}&=&1+\frac{h_ih_j}{4}
\left(\frac{\Delta^2}{\Delta_c^{i}\Delta_L^i\Delta_c^{j}\Delta_L^j}-\frac{1}{\Delta_c^{i}\Delta_L^i}
-\frac{1}{\Delta_c^{j}\Delta_L^j}\right) \nonumber \\
& & + \sum_{l=i,j}\left[\frac{h_l}{2}\left(\frac{\Delta}{\Delta_c^l\Delta_L^l}-
\frac{1}{\Delta}\right)\right.\nonumber\\
& & \left.+\frac{h_l^2}{4}\left(\frac{\Delta^2}{\Delta_c^{l\,2}\Delta_L^{l\,2}}+
\frac{2}{\Delta^2}\right)\right]+\mathcal{O}(h^3) \label{gi}.
\end{eqnarray}
The trace of the product of the perfect time evolution operator
and the one with included hyperfine interaction is
$\mbox{Tr}[U^{\dagger}(0)\,U(h)]=2\{1+\cos[\phi(0)]\cos(\phi)+\cos[(h_i-h_j)t]\sin[\phi(0)]\sin(\phi)\}$.
We obtain for the average fidelity between those two operators to the second order of the Overhauser
fields $h_i$ and $h_j$,
\begin{eqnarray}
F(U)&&=1-\frac{2}{5}\phi^2(0)({\rm D}_i^2 h_i^2+{\rm D}_j^2 h_j^2)\nonumber\\
&&-\frac{2}{5}\sin^2
[\phi(0)]\frac{\phi^2(0)}{\tilde{g}_{ij}^2(0)}(h_i-h_j)^2+\mathcal{O}(h^3),
\label{Fa}
\end{eqnarray}
where ${\rm D}_{l}=(\Delta/\Delta_c^l\Delta_L^l-1/\Delta)/2$, $l=i,j$,
are the first-order coefficients in the Overhauser field in
Eq.~(\ref{gi}), and the time was expressed through the interaction phase $\phi(0)=\tilde{g}_{ij}(0)\, t$.
Averaging of the fidelity (\ref{Fa}) over both Overhauser fields, we find
\begin{eqnarray}
\langle F(U)\rangle_{h_i, h_j} &=&
1-\frac{2}{5}\phi^2(0)\,\sigma^2({\rm D}_i^2+{\rm D}_j^2) \nonumber\\
&&-\frac{4}{5}\frac{\phi^2(0)}{\tilde{g}_{ij}^2(0)}\sin^2[\phi(0)]\sigma^2+\mathcal{O}(h^3).
\label{FH}
\end{eqnarray}
Assuming the parameters
$\Delta_L$ = 4.5 meV, $\Delta_c$ = 5 meV, $g_c$ = 0.5 meV, $\Omega_L^i$ = $\Omega_L^j=$ 1 meV and $\Delta$ = 0.5 meV,
such that ${\rm D}\equiv {\rm D}_i={\rm D}_j=-0.99\times10^{3}\mbox{ eV}^{-1}$and $\tilde{g}_{ij} = 0.02$ meV, the average
fidelity given by Eq.~(\ref{FH}) behaves as function of $\phi(0)$ as shown in Fig.~\ref{FAPlot}.
The value of $\langle F(U)\rangle_{h}$ by $\phi(0)=\pi/4$ is equal to $0.99995$.
As we can see from the formula (\ref{FH}), the
two effects induced by nuclear spins decrease the fidelity with corresponding
strengths: the changing of the electron
spin-spin coupling strength and resonance conditions $\propto {\rm D}^2$
and the precession of the rotation axis of the two-electron spin state $\propto 1/\tilde{g}_{ij}^2(0)$.
These are determined in Eqs. (\ref{geff}), (\ref{gfull}), and (\ref{gi}) and depend on
tunable parameters such as $\Delta$ and $\Delta_L$ (we can express $\Delta_c=\Delta_L+\Delta$). The
cavity detuning can be tuned accordingly, e.\,g. in photonic two-dimentional slab microcavities using Xe condensation \cite{khitrova}.
The contributions to fidelity reduction from terms $\propto {\rm D}^2$ and $\propto 1/\tilde{g}_{ij}^2(0)$ are not of the same order (s. Fig. \ref{FPar} b),
${\rm D}^2\le 0.1/\tilde{g}^2_{ij}(0)$. But we still keep the terms $\propto {\rm D}^2$ in our analytical calculations, because they give the upper boundary
of the fidelity, when $\sin[\phi(0)]=0$. By changing the laser detuning $\Delta_L$ and the detuning $\Delta$ we can
adjust the values ${\rm D}$ and $\tilde{g}_{ij}$ and thus tune the fidelity itself.
The dependence of the fidelity (for $\phi(0)=\tfrac\pi4$) in respect to $\Delta$ and $\Delta_L$ is shown in Fig. \ref{FPar}(a).
It indicates that fidelity can be improved considerably by changing the detunings.
However by choosing better parameters,
some assumptions of the applied theory should be conserved, so that $\tilde{g}_{ij}\ll \Delta$ and $\Delta_L (\Delta_c)\gg g_c$.
One can tune the intensity of applied control laser and rotate the electron spins faster, so that they interact faster as well
and ``feel" the dephasing in a lesser extent. However the coupling strength $\tilde{g}_{ij}(t)$ depends quadratically on laser
intensity and just slight enhancement of Rabi frequency $\Omega_L$ will make it of the same order as $\Delta$.
Therefore, to improve the fidelity and to keep these relations true one can use e. g. $\Delta$ = 0.3 meV and $\Delta_L$ = 4 meV.
With this choise, the average fidelity for the gate $U(\frac\pi4)$ is increased to 0.999991. The fidelity decays considerably slower in this case as well (Fig. \ref{FAPlot}).
\begin{figure}
\includegraphics[width=0.45\textwidth]{analyticalFU}
\caption{Fidelity $F$ of the unitary time evolution operator $U$ generated by the
XY interaction in the presence of nuclear spins as a function of the interaction phase
$\phi(0)\propto t$, where $t$ denotes the interaction time.
Both lines show the analytical result calculated in second order perturbation
theory in the nuclear field. The fidelity can be increased by tuning the parameters $\Delta$ and $\Delta_L$. For smaller $\Delta$ and $\Delta_L$, the fidelity increases due to faster gate operation. The other parameters are held fixed and are given in the text. }
\label{FAPlot}
\end{figure}
\begin{figure}
\includegraphics[width=0.45\textwidth]{FidelityParametric}
\caption{a) The average fidelity $F$ of the unitary time evolution operator $U$ for $\phi(0)=\pi/4$ calculated in second order of the
nuclear field as function of the laser
detuning $\Delta_L$ and two-photon detuning $\Delta$.
An adjustment of these parameters can lead to an increase of the fidelity or to its reduction.
However, the choice of the optimal
parameters is limited by the conditions of the applied formalism.
Further details are given in the text
b) The ratio between two contributing mechanism to reduction of the fidelity ${\rm D}^2(\Delta, \Delta_L)\cdot \tilde{g}^2_{ij}(\Delta, \Delta_L)$. It increases for smaller detunings and consequently for improved fidelities. }
\label{FPar}
\end{figure}
The formula Eq.~(\ref{generalF}) can be applied for calculating the average fidelity of the CNOT operation.
If we neglect the imperfections of the single qubit operations in the CNOT operator sequence
and focus only on the decoherence due to the interaction part, then
we find for the trace distance between a perfect CNOT operator and CNOT
operator with nuclear spins,
\begin{eqnarray}
\mbox{Tr}&&[U^{\dagger}_{\rm CNOT}\,U_{\rm CNOT}(h)]=\\
&&2\left(1+\cos\{2[\phi-\phi(0)]\}\cos\Bigg[(h_i-h_j)\frac{\phi(0)}{\tilde{g}_{ij}(0)}\Bigg]\right)\big|_{\phi(0)=\tfrac\pi4}.\nonumber
\end{eqnarray}
The CNOT average fidelity in second order in the Overhauser field is then
\begin{eqnarray}
F^{\rm CNOT}&=&1-\frac{16}{5} \Big[ \phi(0)=\tfrac\pi4\Big]^2{\rm D}^2\sigma^2 \nonumber\\
&&-\frac{4}{5}\Big[\phi(0)=\tfrac\pi4\Big]^2\frac{\sigma^2}{\tilde{g}^2_{ij}(0)}+ \mathcal{O}(h^3).
\label{FCNOT}
\end{eqnarray}
We find that $F^{\rm CNOT} = 0.9999$ by using $\Delta$ = 0.5 meV and $\Delta_L$ = 4.5 meV, and
$F^{\rm CNOT} = 0.999983$ with $\Delta$ = 0.3 meV and $\Delta_L$ = 4 meV,
which is smaller than the fidelity of
a single $U(\tfrac\pi4)$ operation.
For the average fidelity for further applied CNOT operations, we find
\begin{equation}
F^{\rm CNOT}(n)=\frac{4+|\mbox{Tr}\{(U^{\dagger}_{\rm CNOT})^n\,[U_{CNOT}(h)]^n\}|^2}{20}.
\label{FN}
\end{equation}
The fidelity $F^{\rm CNOT}(n)$ averaged over the nuclear field (Fig.~\ref{FCPlot})
is reduced after each application and can be increased by a better choice of the detuning parameters.
Thus for the last set of parameters only after seven CNOTs, the fidelity drops below typical error correction thresholds ($<$ 0.9999) \cite{nielsen&chuang}.
\begin{figure}
\includegraphics[width=0.48\textwidth]{FCNOT2}
\caption{The analytical and numerical results for fidelity of $n$ subsequent CNOT operations in the presence of
nuclear spins for different detunings. The analytically calculated values are represented by circles and the numerically
calculated ones by stars.}
\label{FCPlot}
\end{figure}
\section{Numerically exact fidelities}
For large values of the parameter $\phi(0)$ we evaluate the averaging of the fidelities over nuclear fields
without expanding $|\mbox{Tr}(U(0)^{\dagger}U(h))|^2$.
\begin{figure}
\includegraphics[width=0.45\textwidth]{numerical_resultsFidUlong}
\caption{Numerically calculated average fidelity of the time evolution
operator for large $\phi(0)$. (a) The comparison between analytical and numerical calculations for $\Delta_L$ = 4.5 meV
and $\Delta$ = 0.3 meV. The analytically obtained fidelity diverges for large $\phi(0)$ while the numerical fidelity saturates to 0.5.
(b) Comparison between two numerically calculated fidelities for the gate $U(\phi)$ with different detunings.}
\label{FullPlot}
\end{figure}
The results of the numerical averaging are shown in the Fig.~\ref{FullPlot}.
The comparison to the analytical result in Fig.~\ref{FullPlot}(a) reveals that our
analytical model can predict the behavior of the fidelity only up to a certain value
of $\phi(0)$. From Fig.~\ref{FullPlot}(a) one can see that around $\phi(0)\simeq$ 12 the
numerical and analytical data begin to differ. For large $\phi(0)$, the analytical data even
diverges into negative values. The numerical data on the other hand converges to 1/2.
The numerical evaluation of the fidelity for different sets of detunings [Fig.~\ref{FullPlot} (b)] confirms
that the tuning of the transition parameters can improve the fidelity significantly. The numerical curve oscillates with
the same period $2\pi$ as the analytical one. The maxima correspond to the situation, when
the fidelity is decreased only due to the modified interaction phase between the electron spin and there is no effect
of the changed rotation axes of the two-electron spin state. The maxima correspond to the decay of the fidelity, where
both Overhauser field induced effects are contributing and the effect of the precession of the rotation axis is maximal.
The fidelity curve suggests that the behavior of its lower boundary
could be described by the Gaussian $\sim\mathrm{e}^{-\phi(0)^2}$.
The fidelity envelope to the second order of $\phi(0)$ is obtained from Eq.~(\ref{FH}) by
setting $\sin (\phi(0))^2$ = 1,
We fit this envelope to
the decay function $\tfrac12[1+\exp(-\phi(0)^2/\tau^2)]$, where $\tau$ is given
through parameters $\sigma$, ${\rm D}$ and $\tilde{g}_{ij}$ from Eq.~(\ref{FH})
and is defined as
\begin{equation}
\tau=\sqrt{\frac{5}{8}}\frac{\tilde{g}_{ij}(0)}{\sigma\sqrt{1+{\rm D}^2\,\tilde{g}^2_{ij}(0)}}
=\tau(\Delta, \Delta_L).
\end{equation}
The Gaussian function gives a good fit to the minimum points of the fidelity for small $\phi(0)$, but
decays faster for larger $\phi(0)$. The tail of the fidelity can be fitted by
adding a third-order term to the
argument of the exponential function,
$\tfrac12[1+\exp(-\phi(0)^2/\tau^2+c\,\phi(0)^3)]$ with $c$ as a fitting parameter.
The fits are shown in Fig.~(\ref{FullPlot}) and fit with $\tau_1$ and $c_1$ corresponds
to the curve with parameters $\Delta_L$ = 4.5 meV and with $\tau_2$ and $c_2$ to numerical curve
with another set of parameters.
The fits yield $c_1 = 9.6\times10^{-7}$ and $c_2 = 5\times10^{-8}$.
The time dependance of the fidelity can be found by changing from
interaction phase $\phi(0)$ to time by assuming $\phi(0)=\tilde{g}_{ij}(0)\,t$.
Then for short time dependance of the lower boundary of the fidelity is described by
the function $0.5[1+\exp(-t^2/T^2)]$, where
\begin{equation}
T= \sqrt{\frac{5}{8}}\frac{\hbar}{\sigma\sqrt{1+{\rm D}^2\,\tilde{g}_{ij}^2(0)}}.
\end{equation}
Since from Fig.~(\ref{FPar}) we know that ${\rm D}^2\,\tilde{g}_{ij}^2(0) \ll 1$, we can neglect it here. It follows that
$T=\sqrt{\frac{5}{8}}\frac{\hbar}{\sigma}=\sqrt{\frac{5}{8}}\frac{\hbar\sqrt{N}}{A}$, where $N$ is the number of the nuclear spins
and $A$ is hyperfine interaction constant.
This implies that the decay of the fidelity does not dependend on the interaction
strength of the electron spins and is only affected by the nuclear field distributions. By tuning the transition parameters
we increase the interaction strength between the coupled electron spins to some reasonable value. Thus the spins acquire
a larger interaction phase within the same time. But the decay time of the time evolution operator fidelity is the same and for
GaAs ($N = 10^5$, $A$ = 90 $\mu$eV) $T=$ 2 ns. The $T_2$ decay time of a electron spin coupled to the nuclear bath is given by
$T_2=\frac{2\hbar\sqrt{N}}{A}$ \cite{coish_prb2004} and in GaAs is $\approx$ 5 ns. Since both interacting spins are affected
by the decoherence, the decay of the common interaction phase is faster than of a single one ($T\approx 0.4 \,T_2$).
The numerically exact CNOT fidelities are presented in Fig. \ref{FCPlot}. The second order approximation in
Overhauser field gives a good agreement to the numerical results. Since the CNOT operator is defined for a fixed interaction phase,
there is no oscillation behavior and the fidelity decreases for each further applied CNOT gates.
\section{Conclusions}
We calculated average fidelities for unitary time evolution operator and CNOT operation of two quantum
dots interacting through an off-resonant cavity mode. We obtained results to second order of
hyperfine interaction
analytically and to an arbitrary order by averaging the fidelities over the Overhauser field distributions
numerically. Both approaches are in good agreement for small interaction time durations.
If only the longitudinal component of the Overhauser field in two interacting quantum dots is considered for the hyperfine interaction, then the time evolution of the two-electron spin state changes in two ways (Fig. \ref{SP}),
and both effects contribute differently to the reduction of the fidelity. A prolongation of the decoherence time
of the two-electron state is possible by narrowing the distribution of the Overhauser field \cite{stepanenko_prl2006} [Eq. (\ref{FH}), (\ref{FCNOT})]. The precession of the rotion axis of the two-electron state on the Bloch sphere can be minimized by preparing the the nuclear spin states in each interacting quantum dot in the same certain state and
minimizing the fluctuations of the Overhauser fields \cite{ribeiro_prl2009}. Since we neglected the hyperfine spin-flip terms in our
calculations, the obtained results are only valid for the Zeeman fields $\propto$ 1 T.
\section{Acknowledgments}
We acknowledge funding from the DFG within SFB 767.
|
\section{Introduction}
Since Hawking proved that black holes can radiate thermally~\cite{Hawking1975}, it is believed that the black holes are kinds of a thermal system and have thermodynamic relations among the quantities describing black holes. There have been especially much efforts to derive the temperature and the entropy of the black holes via various methods. While Hawking used the quantum field theory in curved spacetime in his seminal paper~\cite{Hawking1975}, there exist other methods which give the same predictions. For example, the brick wall method proposed by 't Hooft~\cite{tHooft1985}, or the gravitational anomaly method proposed by Robinson and Wilczek\cite{RobinsonWilczek2005}, or more traditional ways, such as avoiding conical singularity in the Euclidean sector\cite{GibbonsHawking1977}, are suggested. Although all these methods and their modified versions have been successful in deriving the temperature or the entropy of certain types of black holes, they are not satisfactory in the sense that they do not reveal the dynamical nature of the radiation process. The background geometry is fixed mostly in the standard methods. In fact, Hawking and Hartle~\cite{Hartle1976} described the radiation as a pair production. According to this scenario, a pair of particles can be created due to the vacuum fluctuation just inside the horizon. During the positive energy particle tunnel through the horizon to the infinity, the negative energy particle remains inside the hole and reduces the mass of the black hole. However, there is no such intuitive derivation in their original work. As a physical process, it is still important to discuss more realistic models for describing the radiation process.
An idea which treat the Hawking radiation as a semi-classical quantum tunneling process has been proposed by Parikh and Wilczek~\cite{Parikh1999}. Over the past decade, there have been much works applying this method to various types of black holes.(See the references in \cite{Umetsu2011}) The essential idea is that the positive energy particle created just inside the horizon can tunnel through the geometric barrier quantum mechanically. And it is observed as the Hawking flux at infinity. In quantum mechanics, the tunneling probability for classically forbidden paths is given by
\beq
\Gamma \simeq e^{-2\textrm{Im} \textit{I}}, \label{tunneling prob}
\eeq
where $I$ is the action of the particle system.
Though the typical wavelength of the radiation is of the order of the size of the black hole, WKB approximation, or the point particle approximation is justified.\cite{Parikh1999} If we assume that the black holes are in the thermal equilibrium state with the surroundings, the tunneling probability can be identified with the regular Boltzmann factor for a particle of energy $\omega$,
\beq
\Gamma \simeq e^{-\beta \omega},
\eeq
where $\beta$ is the inverse temperature.
One can calculate the Hawking temperature by analyzing the imaginary part of the action for an outgoing particle.
We note that the conservation of energy plays a fundamental role in this approach. The total ADM mass is fixed and the mass of the black hole will decrease while the outgoing particle carries a bunch of energy which amounts to the difference between the total energy and the decreased mass of the black hole.\\
Two different methods have been employed to obtain the imaginary part of the action. One is called the null geodesic method used by Parikh and Wilczek~\cite{Parikh1999}. And the other is called the Hamilton-Jacobi method used by Angheben, Nadalini, Vanzo and Zerbini~\cite{Angheben2005}. In the former method, the contribution to the imaginary part of the action comes from two parts. One is the spacial contribution $\oint p_p d\varphi$, where $p_p$ is the canonical momentum for a particle. The other contribution is a temporal one.\cite{temporal} It is noticed that the tunneling amplitude of the form $\exp(-2\textrm{Im} \int p_p d\varphi)$ is not invariant under the canonical transformations.\cite{Chowdhury2006} The correct expression is $\exp(-\textrm{Im}\oint p_p d\varphi)$ which is identical to (\ref{tunneling prob}) in ordinary quantum mechanics. But they can be, in general, different from each other in general relativity. The temporal contribution must be included because the role of the temporal coordinate and the spacial coordinate are exchanged when we across the horizon. Then it is possible to calculate the imaginary part of the action by using the Hamilton's equation and the metric-dependent information of the null geodesic. The Hamilton-Jacobi method is more direct. One can consider the scalar particles emission and the Klein-Gordon equation governing those particles. Other higher spin particles, of course, can be considered because the Hawking radiation can be composed of particles of various spin. Many authors have succeeded in applying fermion tunneling from various black holes to correctly recover Hawking temperatures~\cite{fermions}. We only consider the neutral scalar particle emission in this letter. By using the WKB approximation, one can obtain the Hamilton-Jacobi equation from the Klein-Gordon equation. Then by solving the Hamilton-Jacobi equation, one can calculate the action directly. Although the original derivation~\cite{Angheben2005} did not consider the self-gravitational effects, a prescription for incorporating the higher order effects of the back reaction is provided.\cite{Medved2005} We also use this modified method to relate the back reaction term to the thermodynamic stability.
In this letter, we apply these two methods to the charged black holes in two dimensional dilaton gravity which is originated from the low energy effective theory of the 0A string theory. The non-chiral projection of two dimensional fermionic string theories gives rise to two types of 0 string theories : type 0A and type 0B. These type 0 theories contain bosonic fields only because NS-NS and RR sectors survive under the non-chiral projection. In type 0A string theory, the NS-NS sector contains a graviton, a dilaton, and a tachyon, while RR sector includes two one-form gauge fields~\cite{Douglas2003}\cite{Takayanagi2003}. It is well known that the low energy effective theory of 0A strings admits charged black hole solutions~\cite{Gukov2003}~\cite{Berkovits2001}. We first derive the correct Hawking temperature via the above two methods and obtain the back reaction term which appears naturally in the tunneling framework. By using the thermodynamic relations, we read off the heat capacity from the back reaction term and show that the heat capacity is positive definite. For usual charged or rotating black holes(for example, Reissner-Nordstr\"{o}m or Kerr-Newman) there exist the critical point at which the second order phase transition occurs.\cite{Davies1977} The heat capacity becomes infinite and discontinuous at that point which characterizes the second order phase transition. And the critical point is determined by the ratio between the the mass, the charge, and the angular momentum. Unlike these black holes, 0A black holes do not allow the negative heat capacity in a physically relevant region. The analysis depends on the ensemble we are dealing with. We mainly use the fixed-charge ensemble in which the thermodynamic stability holds.
The organization of the paper is as follows: In section 2, we briefly review the black hole solutions in type 0A string theory. In section 3, we apply the null geodesic method to the black holes to calculate the imaginary part of the action. After comparing the result with the Boltzmann factor for a particle of energy $\omega$, we can reproduce the Hawking temperature at the leading order in $\omega$. And we obtain the self-interaction effect as a result of applying the energy conservation. We also use the Hamilton-Jacobi method to derive the consistent result. In section 4, we read off the heat capacity from the $\omega^2$ term and the thermodynamic stability is discussed. Section 5 ends with conclusions and discussions on the possibility of the phase transition in the fixed-potential ensemble.
\section{Review of black holes in type 0A string theory}
In this section, we briefly review the black hole solutions in the low energy effective theory of the 0A strings.
The low energy effective action of type 0A string theory at the
lowest order in $\a$ is given by~\cite{Douglas2003}
\bea I_{0A} &=& \int d^2 x \sqrt{-g}\,\bigg[\frac{1}{2\kappa^2}~
e^{-2\Phi}\Big(R + 4\nabla_{\mu}\Phi\nabla^{\mu}\Phi +
\frac{8}{\a} -f_{1}(T)(\nabla T)^2 + f_{2}(T)\Big) \nn \\
&&\qquad \qquad \qquad -\frac{2\pi\a}{4}f_3(T)(F^{+})^2 -
\frac{2\pi\a}{4}f_3(-T)(F^{-})^2 - Q_{+}F^{+} - Q_{-}F^{-}
\bigg]\,,\label{0AString0}\eea
where $F_{\pm}$ denote field strengths of two RR gauge fields and $Q_{\pm}$ denote the corresponding charges, respectively.
The theory admits the following linear dilaton geometry as a vacuum solution:
\bea \Phi &=& -k\varphi\,, \nn \\
ds^2 &=& -dt^2 + d\varphi^2\,,
\label{vacuum}\eea
where all other fields vanish.
It is also known that there exist charged black hole solutions in the model~\cite{Gukov2003}\cite{Berkovits2001}.
When the tachyon field $T$ is turned off, RR gauge fields can be
easily solved as
\begin{equation}
F^{+}_{01}=F^{-}_{01}=\frac{Q}{2\pi\a}\,,\qquad T=0~,
\end{equation}
which corresponds to the configuration of the background D0-branes
with charges given by $Q_{\pm}=Q$\,.
One may use this to integrate out RR gauge fields, and obtain the action of the form
\beq I_{0A}' = \int d^2 x \sqrt{-g}\,\bigg[\frac{1}{2\kappa^2}~
e^{-2\Phi}\Big(R + 4\nabla_{\mu}\Phi\nabla^{\mu}\Phi + 4k^2 \Big) +\Lambda
\bigg]\,.\label{0AString1}\eeq
Here we denoted the original cosmological constant as $k^2 =
2/\alpha'$ and a new effective cosmological constant, coming from
the gauge field contributions, as $\Lambda=-Q^2/(2\pi\a)$.
Therefore the low energy effective theory of type 0A string theory
reduces to the two dimensional dilaton gravity with two kinds of
cosmological constants, one of which is related to the charges of
RR gauge fields. The theory admits the charged black hole
solutions in which the dilaton is taken to be proportional to the
spatial coordinate $\varphi$,
\begin{equation}
\Phi = -k\varphi
\end{equation}
while the black hole geometry is of the form
\beq ds^2 = -f(\varphi) dt^2 + \frac{d\varphi^2}{f(\varphi)} \,,\label{metric}
\eeq
with the factor $f(\varphi)$ given by\footnote{We set $2\kappa^2=1$ for brevity}
\beq f(\varphi) = 1 - \frac{1}{2k} e^{-2k\varphi}
\Big(M-\Lambda\varphi\Big) \,.
\eeq
$M$ may be regarded as a mass of the black hole.
The horizon of the black hole, as an implicit function of $M$, is given by,
\beq
e^{-2k\varphi_H}(M - \Lambda \varphi_H) = 2k. \label{horizon}
\eeq
The extremal black hole solutions are obtained by imposing the following two conditions,
\beq
f(\varphi_{ex})=0, \qquad f'(\varphi_{ex})=0
\eeq
From these conditions one can express the position of the horizon and the mass in terms of the charge,
\beq e^{2k\varphi_{ex}} = - \frac{\Lambda}{4k^2}\,, \qquad M_{ex}
= -\frac{\Lambda}{2k}\Big[1-\ln\Big(-\frac{\Lambda}{4k^2}\Big)\Big]
\,. \eeq
Then the extremal black hole geometry can be expressed as
\beq
f(\varphi) = 1 - e^{-2k(\varphi-\varphi_{ex})}\Big(2k(\varphi-\varphi_{ex}) + 1 \Big). \label{extremal}
\eeq
\section{Hawking radiation as tunneling}
In this section, we derive the Hawking temperature and the back reaction effect of the 0A black holes on the basis of tunneling mechanism. Two methods are used in calculating the imaginary part of the action : the null geodesic method and the Hamilton-Jacobi method.
\subsection{Null geodesic method}
In order to describe the `across-horizon' process, it is necessary to adopt the Painlev\'{e} coordinates which have no coordinate singularity at the horizon.
Introduce new time coordinate to the metric (\ref{metric}) by
\beq
d t \rightarrow d t - \frac{1}{f(\varphi)}\sqrt{1-f(\varphi)}d\varphi. \label{Painleve}
\eeq
Then the line element becomes
\beq
ds^2 = - f(\varphi)dt^2 + 2\sqrt{1-f(\varphi)}dtd\varphi + d\varphi^2. \label{metric2}
\eeq
The null geodesic for the metric (\ref{metric2}) is given by
\bea
\dot{\varphi} &=& \pm 1-\sqrt{1-f(\varphi)} \nn \\
&=& \pm1-\frac{1}{\sqrt{2k}}e^{-k\varphi}\sqrt{M-\Lambda \varphi},
\eea
where the upper(lower) sign correspond to the outgoing(ingoing) geodesics. We take the positive sign to consider the outgoing geodesic which is forbidden classically. On the other hand, the ingoing particle can be ignored because the probability amplitude of the ingoing particle is unity so that the net result of the loop integral $\oint p_p d\varphi$ is the same with that of outgoing particle only.\cite{Chowdhury2006} \\
\beq
\exp \Big(-\textrm{Im}\oint p_p d\varphi \Big) = \exp \Big(-\textrm{Im}\int_{\varphi_{in}}^{\varphi_{out}} p_p d\varphi \Big)
\eeq
The imaginary part of the action for an outgoing positive energy particle is given by
\beq
\textrm{Im} \int_{\varphi_{in}}^{\varphi_{out}} p_p d\varphi = \textrm{Im} \int_{\varphi_{in}}^{\varphi_{out}} \int_0^{p_p} d p_p' d\varphi = - \textrm{Im} \int_{\varphi_{in}}^{\varphi_{out}} \int_M^{M-\omega} \frac{d M'}{\dot{\varphi}(\varphi ; M')}d\varphi,
\eeq
where $p_p$ stands for the canonical momentum of an outgoing particle. In canonical theory, the action actually contains the term involving the integration of the Hamiltonian over time, $d I = p_p d\varphi - H_p dt$.($H_p$ is the Hamiltonian of the particle system.) When we consider the closed loop integral which is invariant under the canonical transformations, the spacial part $\textrm{Im} \int p_p d\varphi$ gives twice the Hawking temperature and the temporal part $H_p \textrm{Im} \Delta t$ gives the exact compensation to yield the well-known value for the Hawking temperature. Essentially the temporal contribution is related to the Painlev\'{e} coordinates (\ref{Painleve}) in which the time coordinate has a pole if one integrates across the horizon. The temporal contribution is a generic feature of the tunneling picture.\cite{temporal} In the last equality above, we have used the Hamilton's equation, $\dot{\varphi}=d H_p / d p_p$ at fixed $\varphi$ and used the energy conservation, $M' = M - \omega'$. Note that the eigenvalue of the Hamiltonian $H_p$ is identified with $\omega$.\\
Then the metric component of the geometry into which the outgoing particles tunnel through is
\beq
f(\varphi) = 1 - \frac{1}{2k} e^{-2k\varphi}
\Big(M-\omega-\Lambda\varphi\Big) \,.
\eeq
The spacial contribution to the imaginary part of the action can be obtained by using Feynman's $i \epsilon$ prescription
\bea
\textrm{Im} \int_{\varphi_{in}}^{\varphi_{out}} p_p d\varphi &=& - \textrm{Im} \int_{\varphi_{in}}^{\varphi_{out}} \Big[ P \int \frac{d M'}{\dot{\varphi}(\varphi ; M')} + i \pi \int \delta(\dot{\varphi}(\varphi ; M'))d M' \Big] d\varphi \nn \\
&=& -\pi \int_{\varphi_{in}}^{\varphi_{out}} \int_M^{M-\omega} \delta \Big( 1-\frac{1}{\sqrt{2k}}e^{-k\varphi}\sqrt{M'-\Lambda \varphi} \Big)d M' d\varphi \nn \\
&=& -\pi \int_{\varphi_{in}}^{\varphi_{out}} 4k e^{2k\varphi} d\varphi \nn \\
&=& \frac{\pi}{k} [\omega - \Lambda(\varphi_{in}-\varphi_{out})], \label{ImS}
\eea
where the sign convention is fixed by requiring the positivity of the coefficient in $\omega$, i.e, inverse temperature at the leading order.\\
To proceed further we have to know the full $\omega$ dependence of $\varphi_{in}-\varphi_{out}$, but this is impossible because of the transcendental character of the metric component,
\beq
e^{-2k\varphi_{in}}
(M-\Lambda\varphi_{in} )=2k \, , \quad e^{-2k\varphi_{out}}
(M-\omega-\Lambda\varphi_{out} )=2k. \label{inout}
\eeq
If we perform a series expansion in $\omega$,
\beq
\varphi_{out} = \varphi_{in}(M-\omega) = \varphi_{in}(M) + \Delta \varphi(\omega),
\eeq
where $\Delta \varphi = \Delta \varphi_1 \omega + \Delta \varphi_2 \omega^2 + \cdots $, we get the results from (\ref{inout}) as follows,
\beq
\Delta \varphi_1 = \frac{-1}{\Lambda + 2k(M-\Lambda \varphi_{in})},
\eeq
\beq
\Delta \varphi_2 = \frac{-2k^2(M-\Lambda \varphi_{in})}{[\Lambda + 2k(M-\Lambda \varphi_{in})]^3}.
\eeq
Then the WKB probability amplitude for the classically forbidden trajectory is given by
\bea
\Gamma &=& \exp \Big[-\textrm{Im} \Big(\oint p_p d\varphi - \omega \Delta t^{in} - \omega \Delta t^{out} \Big)\Big] \nn \\
&=& \exp \Big[
\frac{-4\pi (M-\Lambda \varphi_H)}{\Lambda + 2k(M-\Lambda \varphi_H)}\omega + \frac{2\pi k\Lambda(M-\Lambda\varphi_H)}{[\Lambda + 2k(M-\Lambda \varphi_H)]^3}\omega^2 + \mathcal{O}(\omega^3) \Big]. \label{WKBamp}
\eea
We see that the first term proportional to $\omega$ gives the surface gravity of the black hole.
By comparing with the regular Boltzmann factor for a particle of energy $\omega$, $\Gamma = \exp(-\beta \omega)$, the Hawking temperature is given by
\beq
T_H = \frac{k}{2\pi}\Big[ 1+\frac{\Lambda}{4k^2}e^{-2k\varphi_H} \Big]. \label{temperature}
\eeq
And the quadratic term in $\omega$ is the correction by back reaction of the radiation. This self-interaction effect comes from the energy conservation. Its physical meaning will be discussed in section 4.\\
One can obtain the entropy of the black holes by using standard thermodynamic relations
\beq
d M = T_H d S + \Phi d Q + \Omega d J.
\eeq
There is no angular momentum in the 2D black holes we are concerning about. And we let the charge of the black holes be fixed, therefore we ignore the process of so-called the super radiance. Then the entropy of the black hole becomes
\bea
S &=& \int \frac{dM}{T_H (M)} \nn \\
&=& \frac{2\pi}{k} \int \frac{d \varphi_H}{1+\frac{\Lambda}{4k^2}e^{-2k\varphi_H}}\frac{d M}{d \varphi_H} \nn \\
&=& 4\pi e^{2k\varphi_H}. \label{entropy}
\eea
This is in agreement with the previous results in \cite{Davis2004} and \cite{Hur2008} which are obtained by using the on-shell Euclidean action formalism and the Wald's Noether charge method respectively.\\
\subsection{Hamilton-Jacobi method}
We now calculate the imaginary part of the action making use of the Hamilton-Jacobi equation for the emissive particles.~\cite{Angheben2005}
When black holes emanate radiation at certain temperature, there can be particles of various spin with the same temperature.
Here we consider a massive scalar field in the background metric (\ref{metric}) satisfying the Klein-Gordon equation.\\
\beq
\hbar^2 \nabla^2 \Phi - m^2 \Phi = 0.
\eeq
By using the WKB approximation,
\beq
\Phi = \exp(-\frac{i}{\hbar}I),
\eeq
one can obtain the relativistic Hamilton-Jacobi equation with the limit $\hbar \rightarrow 0$, which gives the action as a function of the coordinates.
\beq
g^{\mu\nu}\p_\mu I \p_\nu I + m^2 = 0.
\eeq
For the metric (\ref{metric}), the Hamilton-Jacobi equation becomes
\beq
-\frac{(\p_t I)^2}{f(\varphi)} + f(\varphi)(\p_\varphi I)^2 + m^2 = 0.
\eeq
We seek a solution of the form
\beq
I(t, \varphi) = -\omega t + W(\varphi).
\eeq
Solving for $W(\varphi)$ yields
\beq
W(\varphi) = \int \frac{d \varphi}{f(\varphi)}\sqrt{\omega^2 - m^2 f(\varphi)}.
\eeq
It is emphasized~\cite{Angheben2005} that we need to adopt the proper spatial distance,
\beq
d\sigma^2 = \frac{d \varphi^2}{f(\varphi)},
\eeq
which is coordinate invariant.\\
By taking the near horizon approximation
\beq
f(\varphi) = f'(\varphi_H)(\varphi - \varphi_H) + \cdots,
\eeq
we find that
\beq
\sigma = \frac{2\sqrt{\varphi-\varphi_H}}{\sqrt{f'(\varphi_H)}},
\eeq
where $0 < \sigma < \infty$.\\
Then the spatial part of the action function reads
\bea
W(\varphi) &=& \frac{2}{f'(\varphi_H)}\int \frac{d\sigma}{\sigma}\sqrt{\omega^2 - \frac{\sigma^2}{4}m^2 f'(\varphi_H)^2} \nn \\
&=& \frac{2\pi i \omega}{f'(\varphi_H)} + \textrm{real contribution}.
\eea
Therefore we obtain the same result with the null geodesic method and the Euclidean method.
\beq
T_H = \beta^{-1} = \frac{\omega}{\textrm{Im} \textit{I} } = \frac{k}{2\pi}\Big[ 1+\frac{\Lambda}{4k^2}e^{-2k\varphi_H} \Big].
\eeq
Here, we also included the temporal contribution which is exactly the same with $\textrm{Im}W$.
However, the above method does not include the back reaction effect. We may incorporate this higher order effect by replacing,~\cite{Medved2005}
\beq
W=\frac{2\pi i \omega}{f'(\varphi_H)} \rightarrow W_q = \int_0^\omega \frac{2\pi i d \omega'}{f'(\varphi_H(M-\omega'))}, \label{actionfunction}
\eeq
where the subscript $q$ denotes the corrected action. Here a gradual transition from $M$ to $M-\omega$ is considered in order to make this energy decrease compatible with the uncertainty principle.\\
Because the explicit form of $\varphi_H$ in terms of $M-\omega'$ is unknown(see (\ref{horizon})), we use
\beq
W_q = \int_{\varphi_{in}(M)}^{\varphi_{out}(M-\omega)} \frac{2\pi i \,d \varphi_H}{f'(\varphi_H)}\frac{d \omega'}{d \varphi_H}
\eeq
instead of (\ref{actionfunction}). In this way we obtain the same result with (\ref{ImS}) and therefore
\beq
\textrm{Im}I = \frac{4\pi (M-\Lambda \varphi_H)}{\Lambda + 2k(M-\Lambda \varphi_H)}\omega - \frac{2\pi k\Lambda(M-\Lambda\varphi_H)}{[\Lambda + 2k(M-\Lambda \varphi_H)]^3}\omega^2 + \mathcal{O}(\omega^3).
\eeq
\section{$\omega^2$ term and thermodynamic stability}
In this section, we discuss about the $\omega^2$ term and its relation to the thermodynamic stability.\\
According to \cite{Medved2005}, the corrected action (\ref{actionfunction}) can be written in a Taylor expanded form by,
\beq
W_q = \frac{i}{2}\int_0^\omega \beta(M-\omega')d \omega' = \frac{i}{2}\Big[ \omega \frac{\p S}{\p M} - \frac{\omega^2}{2}\frac{\p^2 S}{\p M^2} + \mathcal{O}(\omega^3)\Big],
\eeq
where we used the 1st law of thermodynamics, $\beta(M) = \frac{\p S}{\p M} = 4\pi/f'(\varphi_H)$.
And the coefficient of the $\omega^2$ term above is related to other thermodynamic quantities in the fixed-charge ensemble as follows,
\bea
\frac{\p^2 S}{\p M^2} &=& \frac{\p}{\p M} \Big( \frac{1}{T(M)} \Big) \nn \\
&=& -\frac{1}{T^2}\frac{\p T}{\p M} \nn \\
&=& -\frac{1}{T^2}\frac{1}{C_Q}
\eea
where $C_Q$ is the heat capacity of the black holes.
\beq
C_Q = - \frac{1}{T^2}\Big( \frac{\p^2 S}{\p M^2} \Big)^{-1}
\eeq
Just by reading off the coefficient of the $\omega^2$ term,
\beq
\frac{\p^2 S}{\p M^2} = \frac{8\pi k \Lambda(M-\Lambda\varphi_H)}{[\Lambda + 2k(M-\Lambda \varphi_H)]^3},
\eeq
we can verify that the heat capacity of 0A black holes are positive definite in a physically relevant region, that is, $T>0$.
\beq
C_Q = -4\pi e^{2k\varphi_H} \Big( 1 + \frac{4k^2}{\Lambda}e^{2k\varphi_H} \Big) > 0. \label{heat capacity of 0A}
\eeq
For usual charged or rotating black holes, there exist so called the Davies point at which the heat capacity is singular and have different signs in both side around the point. For example, Reissner-Nordstr\"{o}m black holes have a Davies point at $Q=\frac{\sqrt{3}}{2}M$~\cite{Davies1977}, and the second order phase transition occurs at the point. Recently, the physical meaning of the Davies critical point within a tunneling framework was reported focusing on the Reissner-Nordstr\"{o}m and Kerr-Newman black holes.~\cite{La2010} Still it is believed that this critical phenomena are generic for any charged or rotating black holes. But we observe that 0A black holes are quite exceptional : there is no second order phase transition in the fixed-charge ensemble.\\
Let us complete the argument with more rigorous thermodynamic language. In the fixed-charge ensemble, the 1st law of thermodynamics can be written as
\beq
d M_Q = T_Q d S + \Phi d Q,
\eeq
by regarding the parameters $S$(the entropy) and $Q$(the RR charge) as a complete set of global quantities for the black hole thermodynamics. Then the fundamental equation is derived from (\ref{horizon}) and (\ref{entropy}) as follows,
\beq
M_Q(S,Q)=\frac{k S}{2\pi} - \frac{k Q^2}{8\pi}\ln \Big( \frac{S}{4\pi} \Big).
\eeq
And the temperature and the RR potential are given by
\bea
T_Q(S,Q) &=& \frac{\p M_Q}{\p S} = \frac{k}{2\pi} - \frac{k Q^2}{8\pi S}, \nn \\
\Phi (S,Q) &=& \frac{\p M_Q}{\p Q} = -\frac{k Q}{4\pi}\ln \Big( \frac{S}{4\pi} \Big). \label{potential}
\eea
To compute the heat capacity, we use $T_Q$ as a function of $S$ with $Q$ held fixed. Then, we find
\beq
C_Q (S,Q) = T \Big( \frac{\p T_Q}{\p S} \Big)_Q^{-1} = \frac{4 S^2}{Q^2} - S. \label{heat capacity}
\eeq
After inserting $S$ into (\ref{heat capacity}), we verify (\ref{heat capacity of 0A}) in a canonical framework.\\
In general, black holes exhibit the negative heat capacity and its physical meaning is clear. Under the thermodynamic perturbation, the black holes with the negative heat capacity is unstable and therefore can not be regarded as a stable equilibrium state. Hence only a microcanonical ensemble is appropriate. But 0A black holes are free of phase transition and thermodynamically stable. In this connection, the instability region is characterized by a non-physical negative temperature. Hence the 0A black holes can be regarded as a stable equilibrium state within a thermal bath and the canonical ensemble can be used.\\
\section{Conclusion and Discussion}
In this paper we explored the thermodynamic properties of the 0A black holes by using the tunneling picture. First, we applied the null-geodesic method and the Hamilton-Jacobi method to calculate the imaginary part of the action for an outgoing emissive particle. We obtained the correct Hawking temperature and the back reaction effect of the radiation. And we found the heat capacity of the 0A black holes from the back reaction effect which is related to the $\omega^2$ term in the action. The heat capacity of the 0A black holes are found to be positive definite in a physically relevant region. Hence we conclude that the 0A black holes are thermodynamically stable and can be described by a canonical ensemble. And there is no second order phase transition in a fixed-charge ensemble. This is remarkable behavior which does not appear in usual charged or rotating black holes.\\
We only treated the fixed-charge ensemble and showed that there is no phase transition. By changing the representation, we observe another facet of the thermodynamic stability. If we adopt the fixed-potential ensemble, i.e., take the Legendre transformation by using (\ref{potential}),
the mass and the temperature of the 0A black holes are given by
\bea
M_\Phi (S,\Phi) &=& \frac{k S}{2\pi} - \frac{2\pi \Phi^2}{k \ln(S/4\pi)}, \nn \\
T_\Phi (S,\Phi) &=& \frac{k}{2\pi} - \frac{2\pi \Phi^2}{k S \ln(S/4\pi)^2}.
\eea
From this, the heat capacity is found to be
\beq
C_\Phi (\varphi_H,\Phi) = T_\Phi \Big( \frac{\p T_\Phi}{\p S} \Big)_\Phi^{-1} = \frac{4k\varphi_H e^{2k\varphi_H} (4k^4 {\varphi_H}^2 e^{2k\varphi_H} - \pi \Phi^2)}{\Phi^2 (1 + k\varphi_H)}. \label{heat capacity fix po}
\eeq
The heat capacity shows the infinite discontinuity at $\varphi_H = -1/k$. This signals that there might be a meaningful critical point in this ensemble. Since the temperature is positive definite in the range of $| \Phi | < \frac{2k}{e\sqrt{\pi}}$, this point may then be a Davies point. This phenomena at the critical point probably can be classified as a second order phase transition. It will be interesting to study the critical phenomena in the fixed-potential ensemble.\\
{\bf Acknowledgments}
The author would like to thank S.~Hyun and J.~Baek for helpful discussions and comments. Especially, useful comments from D.~Singleton are gratefully acknowledged. This work is supported by the National Research Foundation of Korea(NRF) grant funded by the Korea government(MEST) with the
grant number 2009-0074518.
|
\section{Introduction}
In this paper we consider
the following problem posed in the unit-cube domain
$\Omega=(0,1)^3$
\begin{subequations}\label{eq:Lu}
\begin{align}
\label{eq:Lu_a}
L_{\ve{x}}u(\ve{x})
=-\eps\,\laplace_{\ve{x}}u(\ve{x})-\pt_{x_1}(a(\ve{x})\,u(\ve{x}))+b(\ve{x})\,u(\ve{x})
&=f(\ve{x})
\quad \mbox{for }\ve{x}\in\Omega,\\
u(\ve{x})&=0\,\hspace*{1.0cm}\mbox{for }\ve{x}\in\partial\Omega.
\end{align}
\end{subequations}
Here $\eps$ is a small positive parameter, and we assume that the coefficients
$a$ and $b$ are sufficiently smooth ($a,\,b\in
C^\infty(\bar\Omega)$).
We also assume, for some positive constant $\alpha$, that
\begin{gather}\label{assmns}
a(\ve{x})\geq \alpha>0,
\qquad b(\ve{x})-\pt_{x_1}a(\ve{x})\geq 0
\qquad\mbox{for~all~}\ve{x}\in\bar\Omega.
\end{gather}
Under these assumptions, \eqref{eq:Lu_a} is a singularly
perturbed elliptic equation, also referred to as a
convection-dominated convection-diffusion equation.
Its solutions typically exhibits sharp interior and boundary layers.
This equation
serves as a model for Navier-Stokes equations at large Reynolds
numbers or (in the linearised case) of Oseen equations and
provides an excellent paradigm for numerical techniques in the
computational fluid dynamics \cite{RST08}.
The Green's function for the convection-diffusion problem~\eqref{eq:Lu}
exhibits a strong anisotropic structure, which is demonstrated by Figure~\ref{fig:green}.
This reflects the complexity of solutions
of this problem; it should be noted that
problems of this type
require an intricate asymptotic analysis
\cite[Section IV.1]{Ilin}, \cite{KSh87}; see also \cite[Chapter~IV]{Shi92},
\cite[Chapter~III.1]{RST08} and \cite{KSt05,KSt07}.
We also refer the reader to
D\"orfler \cite{Dorf99}, who, for a similar problem,
gives extensive a priori solution estimates.
\begin{figure}[tb]
\centerline{
\includegraphics[width=0.45\textwidth]{Green_v3}
\includegraphics[width=0.45\textwidth]{Green_2d_version_2}}
\caption{Anisotropy of the Green's function $G$ associated with
\eqref{eq:Lu} for $\eps=0.01$ and $\ve{x}=(\frac15,\frac12,\frac13)$.
Left:
isosurfaces at values of
$1,\,4,\,8,\,16,\,32,\,64,\,128$, and $256$.
Right: a two-dimensional graph for fixed $\xi_3=x_3$.}
\label{fig:green}
\end{figure}
Our interest in considering the Green's function of problem
\eqref{eq:Lu} and estimating its derivatives is motivated by the
numerical analysis of this computationally challenging problem.
More specifically, we shall use the obtained estimates in the
forthcoming paper \cite{FK10_NA} to derive robust a posteriori
error bounds for computed solutions of this problem using
finite-difference methods. (This approach is related to recent
articles \cite{Kopt08,CK09}, which address the numerical solution
of singularly perturbed equations of reaction-diffusion type.)
In a more general numerical-analysis context,
we note that sharp estimates for continuous Green's functions (or
their generalised versions) frequently play a crucial role in a
priori and a posteriori error analyses \cite{erikss,Leyk,notch}.
The purpose of the present paper is to establish
sharp bounds for the derivatives of the Green's function in the
$L_1$ norm (as they will be used to estimate the error in the
computed solution in the dual $L_\infty$ norm \cite{FK10_NA}).
Our estimates will be \emph{uniform in the small perturbation
parameter $\eps$} in the sense that any dependence on $\eps$ will
be shown explicitly.
Note also that our estimates will be {\it sharp}
(in the sense of Theorem~\ref{theo_lower})
up to an $\eps$-independent constant multiplier.
We employ the analysis technique used in
\cite{FK10_1},
which we now extend to a three-dimensional problem. Roughly
speaking, we freeze the coefficients and estimate the
corresponding explicit frozen-coefficient Green's function, and
then we investigate the difference between the original and the
frozen-coefficient Green's functions.
This procedure is often called the parametrix
method.
To make this paper more readable, we deliberately follow some of
the notation and presentation of \cite{FK10_1}.
The paper is organised as follows.
In Section~\ref{sec:def}, the Green's
function associated with problem \eqref{eq:Lu} is defined and
upper bounds for its derivatives are stated in Theorem~\ref{thm:main},
the main result of the paper.
The corresponding lower bounds are then given in Theorem~\ref{theo_lower}.
In Section~\ref{sec:def_green_const}, we obtain the fundamental solution
for a constant-coefficient version of \eqref{eq:Lu} in the domain $\Omega=\R^3$.
This fundamental solution is bounded in
Section~\ref{sec:bounds_green_general}.
It is then used in Section~\ref{sec:bounds_green_const}
to construct certain
approximations of the frozen-coefficient Green's functions
for the domains $\Omega=(0,1)\times\R^2$ and $\Omega=(0,1)^3$.
The difference between these approximations and the original
variable-coefficient Green's function
is estimated in Section~\ref{sec:main_proof}, which completes the proof of
Theorem~\ref{thm:main}.
\medskip
\noindent\textit{Notation.} Throughout the paper, $C$, as well as $c$, denotes
a generic positive constant
that may take different values in different formulas,
but is {\it independent of the small diffusion coefficient $\eps$}.
A subscripted $C$ (e.g., $C_1$) denotes a positive constant
that takes a fixed value, and is also independent of $\eps$.
The usual Sobolev spaces $W^{m,p}(D)$ and $L_p(D)$ on any measurable
domain $D\subset\R^3$ are used. The $L_p(D)$-norm is denoted by $\norm{\cdot}{p\,;D}$
while the $W^{m,p}(D)$-norm is denoted by $\norm{\cdot}{m,p\,;D}$.
By $\ve{x}=(x_1,x_2,x_3)$ we denote an element in $\R^3$.
For an open ball centred at $\ve{x'}$ of radius $\rho$, we use the notation
$B(\ve{x'},\rho)=\{\ve{x}\in\R^3: \sum_{k=1,2,3}(x_k-x'_k)^2<\rho^2\}$.
The notation $\pt_{x_m} f$, $\pt^2_{x_m} f$ and $\laplace_{\ve{x}}$
is employed for the
first- and second-order partial derivatives
of a function $f$ in variable $x_m$, and
the Laplacian in variable $\ve{x}$, respectively, while
$\pt^2_{x_k x_m} f$ will denote a mixed derivative of $f$.
\section{Definition of Green's function. Main result}\label{sec:def}
The Green's function $G=G(\ve{x};\ve\xi)$
associated with \eqref{eq:Lu}, satisfies, for each fixed $\ve{x}\in\Omega$,
\begin{subequations}\label{eq:Green_adj}
\begin{align}
\hspace{-0.3cm}
L^*_{\ve\xi}G(\ve{x};\ve\xi)
:=-\eps\,\laplace_{\ve\xi}G
+a(\ve\xi)\,\pt_{\xi_1}G
+b(\ve\xi)\,G
&=\delta(\ve{x}-\ve\xi)
&&\mbox{for}\;\ve\xi\in\Omega,\\
G(\ve{x};\ve\xi)&=0&&\mbox{for}\;\ve\xi\in\partial\Omega.
\end{align}
\end{subequations}
Here $L^*_{\ve\xi}$ is the adjoint differential operator to $L_{\ve{x}}$,
and $\delta(\cdot)$ is the three-dimensional Dirac $\delta$-distribution.
The unique solution $u$ of \eqref{eq:Lu} allows the representation
\begin{gather}\label{eq:sol_prim}
u(\ve{x})=\iiint_{\Omega}G(\ve{x};\ve\xi)\,f(\ve\xi)\,d\ve\xi\,.
\end{gather}
It should be noted that the Green's function $G$ also satisfies, for each fixed $\ve\xi\in\Omega$,
\begin{subequations}\label{eq:Green_prim}
\begin{align}
L_{\ve{x}}G(\ve{x};\ve\xi)
=-\eps\,\laplace_{\ve{x}}G
-\pt_{x_1}(a(\ve{x})\,G)
+b(\ve{x})\,G
&=\delta(\ve{x}-\ve\xi)&&\mbox{for}\; \ve{x}\in\Omega,\\
G(\ve{x};\ve\xi)&=0&&\mbox{for}\;\ve{x}\in\partial\Omega.
\end{align}
\end{subequations}
Consequently, the unique solution $v$ of the adjoint problem
\begin{subequations}\label{eq:Lu_adj}
\begin{align}
L^*_{\ve{x}}v(\ve{x})=-\eps\,\laplace_{\ve{x}}v
+b(\ve{x})\,\pt_{x_1}v+c(\ve{x})\,v&=f(\ve{x})
\quad \mbox{for}\;\ve{x}\in\Omega,\\
v(\ve{x})&=0\hspace*{1.0cm}\,\mbox{for}\;\ve{x}\in\partial\Omega,
\end{align}
\end{subequations}
is given by
\begin{gather}\label{eq:sol_adj}
v(\ve\xi)=\iiint_{\Omega}G(\ve{x};\ve\xi)\,f(\ve{x})\,d\ve{x}\,.
\end{gather}
We start with a preliminary result for $G$.
\begin{lem}\label{lem_G_L1}
Under assumptions \eqref{assmns}, the Green's function $G$
associated with \eqref{eq:Lu} satisfies
\begin{gather}\label{G_L1}
\iint_{(0,1)^2}\! |G(\ve{x};\ve{\xi})|\,d\xi_2\, d\xi_3
\leq C
,\qquad\quad
\|G(\ve{x};\cdot)\|_{1\,;\Omega}
\leq C,
\end{gather}
where $C$ is some positive $\eps$-independent constant.
\end{lem}
\begin{proof}
The first estimate of \eqref{G_L1} is given in the proof of
\cite[Theorem 2.10]{Dorf99} (see also \cite[Theorem~III.1.22]{RST08}
and \cite{And_2003} for similar two-dimensional results).
The second desired estimate follows.
\end{proof}
We now state the main result of this paper.
\begin{thm}\label{thm:main}
The Green's function $G$ associated with \eqref{eq:Lu},\,\eqref{assmns}
in the unit-cube domain $\Omega=(0,1)^3$ satisfies, for all $\ve{x}\in\Omega$,
the following bounds
\begin{subequations}
\begin{align}
\norm{\pt_{\xi_1} G(\ve{x};\cdot)}{1;\Omega}
&\leq C(1+|\ln \eps|),\label{eq:thm:G_xi}\\
\norm{\pt_{\xi_k} G(\ve{x};\cdot)}{1;\Omega}+
\norm{\pt_{x_k} G(\ve{x};\cdot)}{1;\Omega}
&\leq C\eps^{-1/2},\quad k=2,\,3.\label{eq:thm:G_eta}
\end{align}
Furthermore, for any ball $B(\ve{x}',\rho)$ of radius $\rho$ centered at
any $\ve{x}'\in\bar\Omega$, we have
\begin{align}
\norm{G(\ve{x};\cdot)}{1,1;B(\ve{x}',\rho)}
&\leq C\eps^{-1}\rho,\label{eq:thm:G_grad}
\end{align}
while for the ball $B(\ve{x},\rho)$ of radius $\rho$ centered at $\ve{x}$
we have
\begin{align}
\norm{\pt^2_{\xi_1} G(\ve{x};\cdot)}{1;\Omega\setminus B(\ve{x},\rho)}
&\leq C\eps^{-1}\ln(2+\eps/\rho),\label{eq:thm:G_xixi}\\
\norm{\pt^2_{\xi_k} G(\ve{x};\cdot)}{1;\Omega\setminus B(\ve{x},\rho)}
&\leq C\eps^{-1}(|\ln\eps|+\ln(2+\eps/\rho)),
\quad k=2,\,3.\label{eq:thm:G_etaeta}
\end{align}
\end{subequations}
Here $C$ is some positive $\eps$-independent constant.
\end{thm}
We devote the rest of the paper to the proof of this theorem, which will be completed in
Section~\ref{sec:main_proof}.
In view of the solution representation \eqref{eq:sol_prim},
Theorem~\ref{thm:main} yields
a number of a priori solution estimates for our original problem.
E.g., the bounds
\eqref{eq:thm:G_xi}, \eqref{eq:thm:G_eta} immediately imply the
following result.
\begin{cor}\label{cor_apriori}
Let $f(\ve{x})=\pt_{x_1} F_1(\ve{x})+\pt_{x_2} F_2(\ve{x})+\pt_{x_3} F_3(\ve{x})$
with $F_1,\,F_2,\,F_3\in L_\infty(\Omega)$.
Then there exists a unique solution $u\in L_\infty(\Omega)$
of problem \eqref{eq:Lu}, \eqref{assmns}, for which
we have the bound
\begin{gather}\label{apriori}
\norm{u}{\infty\,;\Omega}
\leq C\bigl[\,(1+|\ln\eps|)\,\norm{F_1}{\infty\,;\Omega}
+ \eps^{-1/2}\,(\norm{F_2}{\infty\,;\Omega}+
\norm{F_3}{\infty\,;\Omega})\,\bigr].
\end{gather}
\end{cor}
It can be anticipated from an
inspection of the bounds for an explicit fundamental solution in a
constant-coefficient case (see Section~\ref{sec:bounds_green_general})
that the upper estimates of Theorem~\ref{thm:main} are {\it sharp}.
Indeed, one can prove the following result.
\begin{thm}[\cite{FK09_lower}]\label{theo_lower}
Let $\eps\in(0,c_0]$ for some sufficiently small positive $c_0$.
Set $a(\ve{x}):=\alpha$ and $b(\ve{x}):=0$ in \eqref{eq:Lu}.
Then the Green's function $G$ associated with this problem
in the unit cube $\Omega=(0,1)^3$ satisfies,
for all $\ve{x}\in[\frac14,\frac34]^3$,
the following lower bounds:
\begin{subequations}\label{eq_thm:main_lower}
\begin{align}
\norm{\pt_{\xi_1} G(\ve{x};\cdot)}{1;\Omega}
&\geq c\,|\ln \eps|,\\
\norm{\pt_{\xi_k} G(\ve{x};\cdot)}{1;\Omega}
&\geq c\,\eps^{-1/2},\,k=2,\,3.
\intertext{Furthermore, for any ball $B(\ve{x};\rho)$ of radius $\rho\le\frac18$, we have}
\norm{G(\ve{x};\cdot)}{1,1;\Omega\cap B(\ve{x};\rho)}
&\geq \begin{cases}
c\,\rho/\eps, & \mbox{for~}\rho\le 2\eps,\\
c\,(\rho/\eps)^{1/2},&\mbox{otherwise},\\
\end{cases}\\
\norm{\pt^2_{\xi_1} G(\ve{x};\cdot)}{1;\Omega\setminus B(\ve{x};\rho)}
&\geq c\,\eps^{-1}\ln(2+\eps/\rho),
\quad&&\mbox{for~}\rho\le c_1\eps,
\\
\norm{\pt^2_{\xi_k} G(\ve{x};\cdot)}{1;\Omega\setminus B(\ve{x};\rho)}
&\geq c\,\eps^{-1}(\ln(2+\eps/\rho)+|\ln\eps|)
&&\mbox{for~}\rho\le{\textstyle\frac18},\,k=2,\,3.
\end{align}
Here $c$ and $c_1$ are $\eps$-independent positive constants.
\end{subequations}
\end{thm}
\section{Fundamental solution in the constant-coefficient case}\label{sec:def_green_const}
In our analysis, we invoke the observation that constant-coefficient versions of
the two problems \eqref{eq:Green_adj} and \eqref{eq:Green_prim}
that we have for $G$, can be easily solved explicitly
when posed in $\R^3$.
So in this section we shall explicitly solve simplifications
of \eqref{eq:Green_adj} and \eqref{eq:Green_prim}. To get these simplifications,
we employ the parametrix method and so
freeze the coefficients in these problems by
replacing $a(\ve{\xi})$ by $a(\ve{x})$ in \eqref{eq:Green_adj}, and
replacing $a(\ve{x})$ by $a(\ve{\xi})$ in \eqref{eq:Green_prim},
and also setting $b:=0$; the frozen-coefficient versions of
the operators $ L^*_{\ve{\xi}}$ and $ L_{\ve{x}}$
will be denoted by $ \bar L^*_{\ve{\xi}}$ and $\tilde L_{\ve{x}}$, respectively.
Furthermore, we extend the resulting
equations to $\R^3$ and denote their solutions by $\bar g$ and~$\tilde g$.
So we get
\begin{align}
\bar L^*_{\ve\xi}\,\bar g(\ve{x};\ve\xi)
=-\eps\,\laplace_{\ve\xi}\bar g(\ve{x};\ve\xi)
+a(\ve{x})\,\pt_{\xi_1}\bar g(\ve{x};\ve\xi)
&=\delta(\ve{x}-\ve\xi)\quad \mbox{for}\;\ve\xi\in\R^3,\label{eq:Green_adj_const}\\
\widetilde L_{\ve{x}}\,\tilde g(\ve{x};\ve\xi)
=-\eps\,\laplace_{\ve{x}}\tilde g(\ve{x};\ve\xi)
-a(\ve\xi)\,\pt_{x_1}\tilde g(\ve{x};\ve\xi)
&=\delta(\ve{x}-\ve\xi)\quad\mbox{for}\; \ve{x}\in\R^3.\label{eq:Green_prim_const}
\end{align}
As $\ve{x}$ appears in \eqref{eq:Green_adj_const} as a parameter, so
the coefficient $a(\ve{x})$ in this equation is considered constant
and we can solve the problem explicitly.
Setting $q=\frac{1}{2}a(\ve{x})$ for fixed $\ve{x}\in(0,1)^3$
and $\bar g(\ve{x};\ve{\xi}) = V(\ve{x};\ve{\xi})\,\E^{q\xi_1/\eps}$ (see, e.g., \cite{KSh87}),
one gets
\begin{align*}
-\eps^2\laplace_{\ve\xi}V+q^2 V
&=\eps \,\E^{-q\xi_1/\eps}\,\delta(\ve{x}-\ve\xi)
=\eps \,\E^{-qx_1/\eps}\,\delta(\ve{x}-\ve\xi).
\end{align*}
As the fundamental solution for the operator
$-\eps^2\laplace_{\ve\xi}+q^2$
is $\frac{1}{4\pi\eps^2} \frac{\E^{-qr/\eps}}{r}$
\cite[Chapter~VII]{TikhSamars},
so
\[
V(\ve{x};\ve{\xi})
=\eps \E^{-x_1q/\eps}\frac{1}{4\pi\eps^2}\frac{\E^{-rq/\eps}}{r}
\quad\mbox{where}\quad
r =\sqrt{(x_1-\xi_1)^2+(x_2-\xi_2)^2+(x_3-\xi_3)^2}.
\]
Finally, for the solution of \eqref{eq:Green_adj_const} we get
\[
\bar g(\ve{x};\ve\xi)
=\frac{1}{4\pi\eps^2}\,\frac{\E^{q(\xi_1-x_1-r)/\eps}}{r},
\qquad\mbox{where}\quad
q=q(\ve{x})={\textstyle\frac{1}{2}}a(\ve{x}).
\]
A similar argument yields the solution of \eqref{eq:Green_prim_const}
\[
\tilde g(\ve{x};\ve\xi)
=\frac{1}{4\pi\eps^2}\,\frac{\E^{q(\xi_1-x_1-r)/\eps}}{r},
\qquad\mbox{where}\quad
q=q(\ve\xi)={\textstyle\frac{1}{2}}a(\ve\xi).
\]
Let $\widehat\xi_{1,[x_1]}=(\xi_1-x_1)/\eps$,
$\widehat\xi_2=(\xi_2-x_2)/\eps$, $\widehat\xi_3=(\xi_3-x_3)/\eps$
and $\widehat{r}_{[x_1]}=\sqrt{\widehat\xi_{1,[x_1]}^2+\widehat\xi_2^2+\widehat\xi_3^2}$.
As we shall need bounds for both $\bar g$ and $\tilde g$,
it is convenient to represent them via a more general function
\begin{gather}\label{eq:def_g0}
g=g(\ve{x};\ve\xi;q)
:=\frac{1}{4\pi\eps^2}\, \frac{\E^{q(\widehat\xi_{1,[x_1]}-\widehat r_{[x_1]})}}{\widehat r_{[x_1]}}
\end{gather}
as
\begin{gather}\label{bar_tilde_g_def}
\bar g(\ve{x};\ve\xi)=g(\ve{x};\ve\xi;q)\Bigr|_{q={\textstyle\frac{1}{2}}a(\ve{x})},
\qquad
\tilde g(\ve{x};\ve\xi)=g(\ve{x};\ve\xi;q)\Bigr|_{q={\textstyle\frac{1}{2}}a(\ve\xi)}
\end{gather}
We use the subindex $[x_1]$ in $\widehat\xi_{1,[x_1]}$ and $\widehat r_{[x_1]}$
to highlight their dependence on $x_1$
as in many places $x_1$ will take different values;
but when there is no ambiguity, we shall sometimes
simply write
$\widehat\xi_1$ and $\widehat r$.
\section{Bounds for the fundamental solution $g(\ve{x};\ve{\xi};q)$}\label{sec:bounds_green_general}
Throughout this section we assume that $\Omega=(0,1)\times\R^2$,
but all results remain valid for $\Omega=(0,1)^3$.
Here we derive a number of useful bounds for the fundamental solution $g$
of \eqref{eq:def_g0}
and its derivatives that will be used in Section~\ref{sec:bounds_green_const}.
As in $\bar g$ and $\tilde g$ we set $q=\frac12 a(\ve{x})$ and $q=\frac12 a(\ve{\xi})$,
respectively, so we shall also use, for $k=2,3$, the differential operators
\begin{gather}
D_{\xi_k}:=\pt_{\xi_k}+{\ts\frac12}\pt_{\xi_k} a(\ve{\xi})\cdot\pt_q,
\qquad\quad
D_{x_k}:=\pt_{x_k}+{\ts\frac12}\pt_{x_k} a(\ve{x})\cdot\pt_q.\label{D_ops}
\end{gather}
\begin{lem}\label{lem:g0_bounds}
Let $\ve{x}\in[-1,1]\times\mathbb{R}^2$ and $0<\frac12\alpha\le q\le C$.
Then for the function $g=g(\ve{x};\ve{\xi};q)$ of \eqref{eq:def_g0}
we have the following bounds
\begin{subequations}\label{g_bounds}
\begin{align}
\label{g_L1} \norm{g(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\\
\label{g_xi_L1} \norm{\pt_{\xi_1} g(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C(1+|\ln\eps|),\\
\label{g_eta_L1} \eps^{1/2}\,\norm{\pt_{\xi_k} g(\ve{x};\cdot;q)}{1\,;\Omega}+
\norm{\pt_q g(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\quad k=2,\,3,\\
\label{g_xi_R_L1} \norm{(\eps \widehat r_{[x_1]}\,\pt_{\xi_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\\
\label{g_eta_x_L1} \eps^{1/2}\,\norm{(\eps \widehat r_{[x_1]}\,\pt^2_{\xi_1\xi_k} g)(\ve{x};\cdot;q)}{1\,;\Omega}+
\norm{(\eps \widehat r_{[x_1]}\,\pt^2_{\xi_1 q} g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\quad k=2,\,3,
\intertext{and for any ball $B(\ve{x}';\rho)$ of radius $\rho$ centered at any $\ve{x}'\in[0,1]\times\mathbb{R}^2$, we have}
\label{g_eta_ball} \norm{g(\ve{x};\cdot;q)}{1,1\,;\Omega\cap B(\ve{x}';\rho)}
&\leq C\eps^{-1}\rho,
\end{align}
while for the ball $B(\ve{x};\rho)$ of radius $\rho$ centered at $\ve{x}$, we have
\begin{align}
\norm{\pt^2_{\xi_1} g(\ve{x};\cdot;q)}{1\,;\Omega\setminus B(\ve{x};\rho)}
&\leq C\eps^{-1}\ln(2+\eps/\rho),\label{g_xi2_L1}\\
\norm{\pt^2_{\xi_k} g(\ve{x};\cdot;q)}{1\,;\Omega\setminus B(\ve{x};\rho)}
&\leq C\eps^{-1}(\ln(2+\eps/\rho)+|\ln\eps|),\quad k=2,\,3.\label{g_eta2_L1}
\end{align}
\end{subequations}
Furthermore, one has the bound
\begin{subequations}
\begin{gather}\label{g_x_L1}
\norm{\pt_{x_1} g(\ve{x};\cdot;q)}{1\,;\Omega}
\leq C(1+|\ln\eps|),
\end{gather}
and with the differential operators \eqref{D_ops}, one has, for $k=2,\,3$,
\begin{align}
\norm{D_{\xi_k} g(\ve{x};\cdot;q)}{1\,;\Omega}+
\norm{D_{x_k} g(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C\eps^{-1/2},\label{D_g}\\
\norm{(\eps \widehat r_{[x_1]}\,D_{\xi_k}\pt_{x_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}+
\norm{(\eps \widehat r_{[x_1]}\,D_{x_k}\pt_{\xi_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C\eps^{-1/2}.\label{D_pt_g}
\end{align}
\end{subequations}
\end{lem}
\begin{proof}
First, note that $\grad_{\ve{x}} g=-\grad_{\ve\xi} g$, so
\eqref{g_x_L1} follows from \eqref{g_xi_L1},
\eqref{D_g} follows from \eqref{D_ops},\,\eqref{g_eta_L1},
while \eqref{D_pt_g} follows from \eqref{D_ops},\,\eqref{g_eta_x_L1}.
Thus it suffices to establish the bounds~\eqref{g_bounds}.
Throughout this proof,
whenever $k$ appears in any relation,
it will be understood to be valid for $k=2,3$
(as all the bounds in~\eqref{g_bounds} that involve $k$,
are given for both $k=2,3$).
A calculation shows that the first-order derivatives of $g=g(\ve{x};\ve{\xi};q)$
are given by
\begin{subequations}\label{diff_g}
\begin{align}
\label{g_xi1} \pt_{\xi_1} g
&=\frac{1}{4\pi\eps^3}\,\widehat r^{-2}\,
\Bigl[q(\widehat r-\widehat \xi_1)-\frac{\widehat\xi_1}{\widehat r}\,\Bigr]
\,\,\E^{q(\widehat\xi_1-\widehat r)},\\
\label{g_xi2}
\pt_{\xi_k} g
&=-\frac{1}{4\pi\eps^3}\,
\bigl(q\widehat r+1\bigr)\,\frac{\widehat\xi_k}{\widehat r^3}
\,\,\E^{q(\widehat\xi_1-\widehat r)}
\\
\label{g_q} \pt_q g
&=\frac{1}{4\pi\eps^2}\,
\frac{\widehat\xi_1-\widehat r}{\widehat r}
\,\,\E^{q(\widehat\xi_1-\widehat r)}.
\end{align}
\end{subequations}
Here we used $\partial_{\xi_j}\widehat r=\eps^{-1}\widehat\xi_j/\widehat r\,$
for $j=1,\,2,\,3$.
In a similar manner, but also using
$\pt_{\xi_i}(\widehat\xi_j/\widehat r)
=-\eps^{-1}\widehat\xi_i\widehat\xi_j/\widehat r^3$ with $i\neq j$,
one gets second-order derivatives
\begin{subequations}\label{diff2_g}
\begin{align}
\label{g2_xi1_xi2} \pt^2_{\xi_1\xi_k} g
&=\frac{1}{4\pi\eps^4}\,
\frac{\widehat\xi_k}{\widehat r^3}\,
\Bigl[q^2(\widehat\xi_1-\widehat r)+
q\frac{3\widehat\xi_1-\widehat r}{\widehat r}+
3\frac{\widehat\xi_1}{\widehat r^2}\,\Bigr]
\,\,\E^{q(\widehat\xi_1-\widehat r)},
\\
\label{g2_xi1_q} \pt^2_{\xi_1 q} g
&=\frac{1}{4\pi\eps^3}\,
\widehat r^{-2}\,
\Bigl[-q(\widehat\xi_1-\widehat r)^2+\frac{\widehat r^2-\widehat\xi_1^2}{\widehat r}\,\Bigr]
\,\,\E^{q(\widehat\xi_1-\widehat r)},\\
\label{g2_xi2_xi2} \pt^2_{\xi_k} g
&=\frac{1}{4\pi\eps^4}\,
\widehat r^{-3}\,\Bigl[q^2\widehat\xi_k^2+(q\widehat r+1)
\frac{3\widehat\xi_k^2-\widehat r^2}
{\widehat r^2}\,\Bigr]
\,\,\E^{q(\widehat\xi_1-\widehat r)}.
\intertext{Finally, combining $\pt_{\xi_1}^2 g=-\pt_{\xi_2}^2 g-\pt_{\xi_3}^2 g+\frac{2q}\eps\,\pt_{\xi_1} g$
with \eqref{g_xi1} and \eqref{g2_xi2_xi2} yields}
\label{g2_xi1_xi1} \pt^2_{\xi_1} g
&=\frac{1}{4\pi\eps^4}
\,\widehat r^{-3}\,\Bigl[ q^2\bigl(\widehat r-\widehat\xi_1\bigr)^2
-q \bigl(\widehat r-\widehat\xi_1\bigr)
\Bigl(1+3\frac{\widehat\xi_1}{\widehat r}\Bigr)
+\frac{3\widehat\xi_1^2-\widehat r^2}
{\widehat r^2}\,\Bigr]
\,\,\E^{q(\widehat\xi_1-\widehat r)}.
\end{align}
\end{subequations}
Now we proceed to estimating the above derivatives of $g$.
Note that $d\ve{\xi}=\eps^3 d\widehat{\ve{\xi}}$,
where $\widehat{\ve{\xi}}\in\widehat\Omega:=\eps^{-1}(-x_1,1-x_1)\times\R^2
\subset(-\infty,2/\eps)\times\R^2$.
Consider the two sub-domains
\[
\widehat\Omega_1:=\bigl\{\;\widehat\xi_1<1+{\ts\frac{1}{2}}\widehat r
\;\bigr\},
\qquad
\widehat\Omega_2:=\bigl\{\;
\max\{\,1,{\ts\frac{1}{2}}\widehat r\,\}<\widehat\xi_1<2/\eps\;\bigr\}.
\]
As $\widehat\Omega\subset\widehat\Omega_1\cup\widehat\Omega_2$ for any $x_1\in[-1,1]$,
it is convenient to consider integrals over these two sub-domains separately.
(i) Consider $\widehat{\ve{\xi}}\in\widehat\Omega_1$. Then
$\widehat\xi_1\leq 1+\frac{1}{2}\widehat r$, so
one gets
\begin{align}
\eps^3\bigl[(1+\widehat r)
(\eps^{-1}|g|+|\pt_{\xi_1} g|+|\pt_{\xi_k} g
+|\pt_q g|+|\pt^2_{\xi_1 q} g|)
&+\eps\widehat r|\pt^2_{\xi_1\xi_k} g
\bigr]\notag\\
&\leq C \, \widehat r^{-2}\, (1+\widehat r+\widehat r^2+\widehat r^3)
\,\,\E^{q(\widehat\xi_1-\widehat r)}\notag\\
&\leq C\,\widehat r^{-2}\,\, \E^{-q\widehat r/4},\label{star0}
\end{align}
where we combined $\E^{q\widehat\xi_1}\le \E^{q(1+\widehat r/2)}$
with $(1+\widehat r+\widehat r^2+\widehat r^3)\le C\E^{q\widehat r/4}$.
This immediately yields
\begin{multline}
\iiint_{\widehat\Omega_1}\!
\bigl[(1+\widehat r)
(\eps^{-1}|g|+|\pt_{\xi_1} g|+|\pt_{\xi_k} g
+|\pt_q g|+|\pt^2_{\xi_1 q} g|)
+\eps\widehat r|\pt^2_{\xi_1\xi_k} g
\bigr]\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\\
\leq C\int_0^{\infty}\E^{-q\widehat r/4} \,d\widehat r
\leq C.\label{star}
\end{multline}
Similarly,
\[
\eps^3\bigl[|\pt^2_{\xi_1} g|+|\pt^2_{\xi_k} g
\bigr]
\le C\eps^{-1}\,\widehat r^{-3}\,(1+\widehat r^2)\,\,\E^{q(\widehat\xi_1-\widehat r)}
\le C\eps^{-1}\,\widehat r^{-2}\,(\widehat r^{-1}+\widehat r)\,\,\E^{-q\widehat r/2},
\]
so
\begin{gather}\label{star2}
\iiint_{\widehat\Omega_1\setminus B(\ve{0};\widehat\rho)}\!\bigl[|\pt^2_{\xi_1} g|+|\pt^2_{\xi_k} g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\eps^{-1}\!\int_{\widehat\rho}^{\infty}\!\!(\widehat r^{-1}+\widehat r)\,\E^{-q\widehat r/2} \,d\widehat r
\leq C\eps^{-1}\ln(2+\widehat\rho^{-1}).
\end{gather}
Furthermore, for an arbitrary ball $\widehat B_{\widehat\rho}$ of radius $\widehat\rho$
in the coordinates $\ve{\widehat\xi}$, we get
\begin{gather}\label{ball_Omega1}
\iiint_{\widehat\Omega_1\cap \widehat B_{\widehat\rho}}\!
\bigl[|g|+|\pt_{\xi_1} g|+|\pt_{\xi_k} g
\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\int_0^{\widehat\rho}\!\!\E^{-q\widehat r/4} \,d\widehat r\leq C\min\{\widehat\rho,1\}.
\end{gather}
(ii) Next consider $\widehat{\ve{\xi}}\in\widehat\Omega_2$.
In this sub-domain, it is convenient to rewrite the integrals in terms of
$(\widehat \xi_1,t_2,t_3)$, where
\begin{gather}\label{t_def}
t_k:=\widehat\xi_1^{-1/2}\,\widehat\xi_k,
\quad\mbox{so}\quad
\widehat\xi_1^{-1/2}\,d\widehat\xi_k=dt_k
\quad\mbox{and}\quad
\widehat r-\widehat\xi_1=\frac{\widehat\xi_2^2+\widehat\xi_3^2}{\widehat r+\widehat\xi_1}\le t_2^2+t_3^2=:t^2.
\end{gather}
Note that $\widehat\xi_1\le\widehat r\le 2\,\widehat\xi_1$ in $\widehat\Omega_2$
so $\widehat{r}-\widehat\xi_1=(\widehat\xi_2^2+\widehat\xi_3^2)/(\widehat{r}+\widehat\xi_1)\ge c_0 t^2$,
where $c_0:=\frac1
$.
Consequently $\E^{-q(\widehat{r}-\widehat\xi_1)}\leq \E^{-q
c_0t^2}$ or
\begin{gather}\label{Q_def}
\E^{-q(\widehat{r}-\widehat\xi_1)}\leq C\, \widehat r\,
Q,
\qquad\mbox{where}\quad
Q:=\widehat\xi_1^{-1}\, \E^{-q c_0t^2},
\end{gather}
and
\begin{align}\notag
\iint_{\R^2}(1+t+t^2+t^3+t^4)\,Q\,\,d\widehat\xi_2\,d\widehat\xi_3
&= \iint_{\R^2} (1+t+t^2+t^3+t^4)\,\E^{-qc_0t^2}
\,dt_2\,dt_3
\\\label{Q_int}
&\leq C.
\end{align}
Using \eqref{diff_g},\eqref{diff2_g} and \eqref{t_def} it is straightforward to prove the
following bounds for $g$ and its derivatives in $\widehat\Omega_2$
\begin{subequations}\label{bounds15}
\begin{align}
\label{bound0} \eps^3| g|
&\leq C\,\eps\, Q,\\
\label{bound1} \eps^3|\pt_{\xi_k} g|
&\leq C\,\widehat\xi_1^{-1/2}\, t\, Q
\\
\label{bound2} \eps^3|\pt^2_{\xi_k} g|
&\leq C\,\eps^{-1}\,\widehat\xi_1^{-1}\, [1+t^2]\, Q,
\intertext{and also}
\label{bound3} \eps^3(\eps \widehat r|\pt_{\xi_1} g|+|\pt_q g|)
&\leq C\,\eps\, [1+t^2]\, Q,\\
\label{bound4} \eps^3|\pt_{\xi_1} g|
&\leq C\,\widehat\xi_1^{-1}\, [1+t^2]\, Q,\\
\label{bound5} \eps^3(\eps\widehat r|\pt_{\xi_1\xi_k} g|)
&\leq C\,\widehat\xi_1^{-1/2}\, t\,[1+t^2]\, Q
\\
\label{bound6} \eps^3(\eps \widehat r|\pt^2_{\xi_1 q} g|)
&\leq C\,\eps\, (t^2+t^4)\, Q,\\
\label{bound7} \eps^3|\pt^2_{\xi_1} g|
&\leq C\,\eps^{-1}\, \widehat\xi_1^{-2}\,(1+t^2+t^4)\, Q.
\end{align}
\end{subequations}
Combining the obtained estimates \eqref{bounds15} with \eqref{Q_int} yields
\begin{align}
\hspace{-0.2cm}
\iiint_{\widehat\Omega_2}\!
\bigl[|g|+\eps^{1/2}|\pt_{\xi_k} g
+\eps\widehat r|\pt_{\xi_1} g|+|\pt_q g|
&+\eps^{1/2}\eps\widehat r|\pt^2_{\xi_1\xi_k} g|
+\eps\widehat r|\pt^2_{\xi_1 q} g|
+|\pt_{\xi_1}^2 g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)\notag\\
&\leq C\int_1^{2/\eps}\!\![\eps+\eps^{1/2}\widehat\xi_1^{-1/2}] \,
d\widehat\xi_1
\leq C.\label{int_Omega2_main}
\end{align}
Similarly, combining \eqref{bound2} and \eqref{bound4} with \eqref{Q_int} yields
\begin{gather}\label{int_Omega2_aux}
\iiint_{\widehat\Omega_2}\!\bigl[|\pt_{\xi_1} g
+\eps|\pt^2_{\xi_k} g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\int_1^{2/\eps}\!\!\widehat\xi_1^{-1} \,d\widehat\xi_1
\leq C(1+|\ln\eps|).
\end{gather}
Furthermore, by \eqref{bound1}, and \eqref{bound4}
for an arbitrary ball $\widehat B_{\widehat\rho}$ of radius $\widehat\rho$
in the coordinates $\widehat{\ve{\xi}}$, we get
\begin{gather}\label{ball_Omega2}
\iiint_{\widehat\Omega_2\cap \widehat B_{\widehat\rho}}\!
(|g|+|\pt_{\xi_1} g
+|\pt_{\xi_k} g|)
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\int_1^{1+\widehat\rho}\!\!
\bigl[\eps+\widehat\xi_1^{-1}+\widehat\xi_1^{-1/2}\bigr]
\,d\widehat\xi_1
\leq C\widehat\rho.
\end{gather}
To complete the proof, we now recall that $\widehat\Omega\subset\widehat\Omega_1\cup\widehat\Omega_2$ and
combine estimates \eqref{star} and \eqref{star2} (that involve integration over $\widehat\Omega_1$)
with \eqref{int_Omega2_main} and \eqref{int_Omega2_aux},
which yields the desired bounds \eqref{g_L1}-\eqref{g_eta_x_L1}
and \eqref{g_xi2_L1}, \eqref{g_eta2_L1}.
To get the latter two bound we also used the observation that the ball $B(\ve{x};\rho)$ of radius $\rho$
in the coordinates $\ve{\xi}$ becomes the ball $B(\ve{0};\widehat\rho)$ of radius $\widehat\rho=\eps^{-1}\rho$
in the coordinates $\widehat{\ve{\xi}}$.
The remaining assertion \eqref{g_eta_ball}
is obtained by combining \eqref{ball_Omega1} with \eqref{ball_Omega2} and
noting that an arbitrary ball $B(\ve{x}';\rho)$ of radius $\rho$
in the coordinates $\ve{\xi}$
becomes a ball $\widehat B_{\widehat\rho}$ of radius $\widehat\rho=\eps^{-1}\rho$
in the coordinates $\widehat{\ve{\xi}}$.
\end{proof}
Our next result shows that for $x_1\ge 1$, one gets stronger bounds for $g$ and its derivatives.
These bounds involve the weight function
\begin{gather}\label{lambda_def}
\lambda:=\E^{2q(x_1-1)/\eps}.
\end{gather}
and show that, although $\lambda$ is exponentially large in $\eps$,
this is compensated by the smallness of $g$ and its derivatives.
\begin{lem}\label{lem:g_lmbd_bounds}
Let $\ve{x}\in[1,3]\times\R^2$ and $0<\frac12\alpha\le q\le C$.
Then for the function $g=g(\ve{x};\ve{\xi};q)$ of \eqref{eq:def_g0}
and the weight $\lambda$ of \eqref{lambda_def},
one has the following bounds
\begin{subequations}\label{g_bounds_lmbd_}
\begin{align}
\norm{([1+\eps\widehat r_{[x_1]}]\,\lambda g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C\eps,\label{g_L1_mu}\\
\norm{(\lambda\,\pt_{\xi_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\norm{(\lambda\,\pt_q g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\label{g_L1_mu2}\\
\norm{([1+\eps^{1/2}\widehat r_{[x_1]}]\,\lambda\,\pt_{\xi_k} g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\eps^{1/2}\norm{(\eps\widehat r_{[x_1]}\,\lambda\,\pt_{\xi_1\xi_k}^2 g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\quad k=2,3,
\label{g_L1_mu3}\\
\norm{\widehat r_{[x_1]}\,\pt_q(\lambda\, g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\norm{\eps \widehat r_{[x_1]}\,\pt_q(\lambda\,\pt_{\xi_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\le C,\label{g_L1_mu4}
\intertext{and for any ball $B(\ve{x}';\rho)$ of radius $\rho$ centered at any
$\ve{x}'\in[0,1]\times\R^2$, one has}
\norm{(\lambda\, g)(\ve{x};\cdot;q)}{1,1\,;\Omega\cap B(\ve{x}';\rho)}
&\leq C\eps^{-1}\rho,\label{g_L1_mu5}
\end{align}
while for the ball $B(\ve{x};\rho)$ of radius $\rho$ centered at $\ve{x}$ and $k=1,\,2,\,3$, one has
\begin{gather}\label{g_L1_mu6}
\norm{(\lambda\,\pt^2_{\xi_k} g)(\ve{x};\cdot;q)}{1\,;\Omega\setminus B(\ve{x},\rho)}
\leq C\eps^{-1}\ln(2+\eps/\rho).
\end{gather}
\end{subequations}
Furthermore, with the differential operators \eqref{D_ops} and $k=2,\,3$, we have
\begin{subequations}\label{g_lmb_bounds}
\begin{align}
\norm{\pt_{x_1}(\lambda g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\norm{D_{x_k}(\lambda g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\norm{D_{\xi_k}(\lambda g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C,\label{D_g_lmbd}\\
\norm{\eps \widehat r_{[x_1]}\,D_{x_k}(\lambda\, \pt_{\xi_1} g)(\ve{x};\cdot;q)}{1\,;\Omega}
+\norm{\eps \widehat r_{[x_1]}\,D_{\xi_k}\pt_{x_1}(\lambda g)(\ve{x};\cdot;q)}{1\,;\Omega}
&\leq C\eps^{-1/2}.\label{D_pt_lmbd}
\end{align}
\end{subequations}
\end{lem}
\begin{proof}
Throughout this proof,
whenever $k$ appears in any relation, it will be understood to be
valid for $k=2,3$ (as all the bounds in~\eqref{g_bounds_lmbd_},
\eqref{g_lmb_bounds} that involve $k$,
are given for both $k=2,3$).
We shall use the notation $A=A(x_1):=(x_1-1)/\eps\geq 0$.
Then \eqref{lambda_def} becomes $\lambda=\E^{2qA}$.
We partially imitate the proof of Lemma~\ref{lem:g0_bounds}.
Again $d\ve{\xi}=\eps^3\,d\widehat{\ve{\xi}}$,
but now $\widehat{\ve{\xi}}\in\widehat\Omega=\eps^{-1}(-x_1,1-x_1)\times\R^2
\subset(-3/\eps,-A)\times\R^2$.
So $\widehat\xi_1<-A\le 0$ immediately yields
\begin{gather}\label{lambda_exp}
\lambda\, \E^{q \widehat\xi_1}=\E^{2q(A-|\widehat\xi_1|)}\,\E^{q|\widehat\xi_1|}\le \E^{q|\widehat\xi_1|}.
\end{gather}
Consider the sub-domains
\begin{align*}
\widehat\Omega'_1&:=\bigl\{\;|\widehat\xi_1|<1+{\ts\frac12\widehat r},\;\;\widehat\xi_1<-A\;\bigr\},\\
\widehat\Omega'_2&:=\bigl\{\;|\widehat\xi_1|>\max\{\ts1,{\ts\frac12\widehat r}\},\;\;
-3/\eps<\widehat\xi_1<-A\;\bigr\}.
\end{align*}
As $\widehat\Omega\subset\widehat\Omega_1'\cup\widehat\Omega_2'$
for any $x_1\in[1,3]$, we estimate integrals over these two domains
separately.
(i) Let $\widehat{\ve{\xi}}\in\widehat\Omega'_1$. Then $|\widehat\xi_1|\leq 1+\frac12\widehat r$
so, by \eqref{lambda_exp}, one has $\lambda\, \E^{q \widehat\xi_1}\le \E^{q(1+\widehat r/2)}$.
The first inequality in \eqref{star0} remains valid, but now we combine it with
\begin{gather}\label{eq:lambda_eqxi}
\lambda \,\E^{q(\widehat\xi_1-\widehat r)}\, (1+\widehat r+\widehat r^2+\widehat r^3)\,
\le C\, \E^{-q\widehat r/4}
\end{gather}
(which is obtained similarly to the final line in \eqref{star0}).
This leads to a version of \eqref{star} that involves the weight $\lambda$:
\begin{multline}\label{star_lambda}
\iiint_{\widehat\Omega_1'}\lambda\,
\bigl[(1+\widehat r)
(\eps^{-1}|g|+|\pt_{\xi_1} g|+|\pt_{\xi_k} g
+\eps^{-1}|\pt_q g|+|\pt^2_{\xi_1 q} g|)
+\eps\widehat r|\pt^2_{\xi_1\xi_k} g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C.
\end{multline}
In a similar manner, we obtain versions of estimates \eqref{star2}
and \eqref{ball_Omega1}, that also involve the weight $\lambda$:
\begin{gather}\label{star2_lambda}
\iiint_{\widehat\Omega_1'\setminus B(\ve{0};\widehat\rho)}\!
\lambda\,|\pt^2_{\xi_k} g
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\eps^{-1}\ln( 2+\widehat\rho^{-1}),
\end{gather}
\begin{gather}\label{ball_Omega1_lambda}
\iiint_{\widehat\Omega_1'\cap \widehat B_{\widehat\rho}}\!
\lambda\bigl[|g|+|\pt_{\xi_1} g
+|\pt_{\xi_k} g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
\leq C\min\{\widehat\rho,1\},
\end{gather}
where $\widehat B_{\widehat\rho}$ is an arbitrary ball of radius $\widehat\rho$
in the coordinates $\widehat{\ve{\xi}}$.
Furthermore, \eqref{star_lambda} combined with
$|\pt_q(\lambda\, g)|\le \lambda(2A|g|+|\pt_q g|)$
and $|\pt_q(\lambda\,\pt_{\xi_1} g)|\leq \lambda(2A|\pt_{\xi_1} g|+|\pt^2_{\xi_1 q} g|)$
and then with $A\le 2/\eps$ yields
\begin{gather}\label{star_lambda2}
\iiint_{\widehat\Omega_1'}\, \widehat r\,\bigl[
|\pt_q(\lambda\, g)|+\eps|\pt_q(\lambda\,\pt_{\xi_1} g)|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)\leq C.
\end{gather}
(ii) Now consider $\widehat{\ve{\xi}}\in\widehat\Omega'_2$.
In this sub-domain (similarly to $\widehat\Omega_2$ in the proof of Lemma~\ref{lem:g0_bounds})
one has $|\widehat\xi_1|\le \widehat r\le2|\widehat\xi_1|$ and
$c_0 t^2\le\widehat r-|\widehat\xi_1|\le t^2$, where
$t_k:=|\widehat\xi_1|^{-1/2}\,\widehat\xi_k$ for $k=2,3$,
and
$t^2:=t_2^2+t_3^2$, (compare with \eqref{t_def}).
We also introduce a new barrier $Q$
\begin{gather}\label{Q_def2}
Q:=\lambda^{-1}\,\E^{2q(A-|\widehat\xi_1|)}\,\bigl\{|\widehat\xi_1|^{-1}\, \E^{-q c_0t^2}\bigr\}
\qquad\Rightarrow\quad
\E^{-q(\widehat{r}-\widehat\xi_1)}\leq C\, \widehat r\,
Q,
\end{gather}
(compare with \eqref{Q_def}; to get the bound for $\E^{-q(\widehat{r}-\widehat\xi_1)}$ we used
\eqref{lambda_exp}).
With the new definition \eqref{Q_def2} of $Q$, the bounds \eqref{bound0}--\eqref{bound2}
remain valid in $\widehat\Omega_2'$ only with $\widehat\xi_1$ replaced by $|\widehat\xi_1|$.
Note that the bounds \eqref{bound3}--\eqref{bound6} are not valid in $\widehat\Omega_2'$,
(as they were obtained using $\widehat r-\widehat\xi_1\leq t^2$,
which is not the case for $\widehat\xi_1<0$).
Instead, using $\widehat r\geq |\widehat\xi_1|\geq 1$ and
$\widehat r\leq2|\widehat\xi_1|$,
we prove, directly from \eqref{diff_g},\eqref{diff2_g},
the following bounds in $\widehat\Omega'_2$:
\begin{subequations}\label{bounds_lambda}
\begin{align}
\eps^3|\pt_{\xi_1} g|
&\leq C\, Q, \label{bound2b_lmb}\\
\eps^3|\pt_q g|
&\leq C\,\eps |\widehat\xi_1|\, Q,\label{bound2_lmb}\\
\eps^3(\eps\widehat r|\pt_{\xi_1\xi_k} g|)
&\leq C\,|\widehat\xi_1|^{1/2}\, t\, Q
\label{bound5_lmb}\\
\eps^3(|\pt_q(\lambda\, g)|+\eps|\pt_q(\lambda\,\pt_{\xi_1} g)|)
&\leq C\,\eps \lambda\, [(|\widehat\xi_1|-A)+t^2+1]\, Q.\label{bound2a_lmb}
\end{align}
\end{subequations}
In particular, to establish \eqref{bound2a_lmb}, we combined
$\pt_q(\lambda\, g)=\lambda[2A\,g+\pt_{q}g]$ and
$\pt_q(\lambda\,\pt_{\xi_1} g)=\lambda[2A\,\pt_{\xi_1} g+\pt^2_{\xi_1 q}g]$
with the observations that
\[
(\widehat{r}+|\widehat\xi_1|)-2A
= 2(|\widehat\xi_1|-A)+(\widehat r-|\widehat\xi_1|)
\leq 2(|\widehat\xi_1|-A)+t^2
\]
and $\widehat r^{-1}A\leq C$.
Next, note that \eqref{Q_int} is valid with $Q$
replaced by
the multiplier $\bigl\{|\widehat\xi_1|^{-1}\, \E^{-q c_0t^2}\bigr\}$ from
the current definition \eqref{Q_def2}
of $Q$.
Combining this observation with the bounds \eqref{bound0}--\eqref{bound2} and
\eqref{bound2b_lmb}--\eqref{bound5_lmb}, and also with
$\widehat r\le 2|\widehat\xi_1|$, yields
\begin{align}
\iiint_{\widehat\Omega_2'}\!\! \lambda\,
\bigl[&(\eps^{-1}+\widehat r)|g|+|\pt_{\xi_1} g|
+(1+\eps^{1/2}\widehat r)|\pt_{\xi_k} g|
+|\pt_q g|
+\eps^{1/2}(\eps\widehat r|\pt^2_{\xi_1\xi_k} g|)
+\eps|\pt^2_{\xi_k} g|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)\notag\\
&\leq C\int_{-3/\eps}^{-\max\{A,1\}}\!
\bigl[1+\eps|\widehat\xi_1|+|\widehat\xi_1|^{-1/2}
+(\eps|\widehat\xi_1|)^{1/2}+|\widehat\xi_1|^{-1}\bigr]\,
\E^{2q(A-|\widehat\xi_1|)} \,d\widehat\xi_
\leq C.\label{int_Omega2_main_lmbd}
\end{align}
Similarly, from \eqref{bound2a_lmb} combined with
$\widehat r\leq 2|\widehat\xi_1|\le 6\eps^{-1}$, one gets
\begin{multline}\label{Omega2_prime}
\iiint_{\widehat\Omega_2'}\!
\widehat r\bigl[|\pt_q(\lambda\, g)|+\eps|\pt_q(\lambda\,\pt_{\xi_1} g)|\bigr]
\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)\\
\leq C\int_{-3/\eps}^{-\max\{A,1\}}\!\!
\bigl[(|\widehat\xi_1|-A)+1\bigr]\,\E^{2q(A-|\widehat\xi_1|)} \,d\widehat\xi_1
\leq C.
\end{multline}
Furthermore, by \eqref{bound1}, and \eqref{bound2b_lmb},
for an arbitrary ball $\widehat B_{\widehat\rho}$ of radius $\widehat\rho$
in the coordinates $\widehat{\ve{\xi}}$, we get
\begin{align}\notag
\iiint_{\widehat\Omega'_2\cap \widehat B_{\widehat\rho}}\!
\lambda\,[|g|+|\pt_{\xi_1} g
+|\pt_{\xi_k} g|]\,\bigl(\eps^3d\widehat{\ve{\xi}}\bigr)
&\leq C\int^{-\max\{A,1\}}_{-\max\{A,1\}-\widehat\rho}
\bigl[1+|\widehat\xi_1|^{-1/2}\bigr]\,\E^{2q(A-|\widehat\xi_1|)}d\widehat\xi_1
\\\label{ball_Omega2_lambda}
&\leq C\widehat\rho.
\end{align}
To complete the proof of \eqref{g_bounds_lmbd_}, we now
recall that $\widehat\Omega\subset\widehat\Omega_1'\cup\widehat\Omega_2'$ and
combine estimates \eqref{star_lambda}, \eqref{star2_lambda}, \eqref{star_lambda2}
(that involve integration over $\widehat\Omega_1'$)
with \eqref{int_Omega2_main_lmbd}, \eqref{Omega2_prime},
which yields the desired bounds \eqref{g_L1_mu}--\eqref{g_L1_mu4}
and the bounds for $\pt_{\xi_2}^2g$ and $\pt_{\xi_3}^2 g$ in \eqref{g_L1_mu6}.
To get the latter two bounds we also used the observation
that the ball $B(\ve{x};\rho)$ of radius $\rho$
in the coordinates $\ve{\xi}$
becomes the ball $B(\ve{0};\widehat\rho)$ of radius $\widehat\rho=\eps^{-1}\rho$
in the coordinates $\widehat{\ve{\xi}}$.
The bound for $\pt^2_{\xi_1} g$ in \eqref{g_L1_mu6}
follows as $\pt^2_{\xi_1} g=-\pt^2_{\xi_2} g-\pt^2_{\xi_3} g+\frac{2q}\eps\,\pt_{\xi_1} g$
for $\ve{\xi}\neq\ve{x}$.
The remaining assertion \eqref{g_L1_mu5} is obtained by combining
\eqref{ball_Omega1_lambda} with \eqref{ball_Omega2_lambda} and
noting that an arbitrary ball $B(\ve{x}';\rho)$ of radius $\rho$
in the coordinates $\ve{\xi}$
becomes a ball $\widehat B_{\widehat\rho}$ of radius $\widehat\rho=\eps^{-1}\rho$
in the coordinates $\widehat{\ve{\xi}}$.
Thus we have established all the bounds \eqref{g_bounds_lmbd_}.
We now proceed to the proof of the bounds \eqref{g_lmb_bounds}.
Note that $\grad_{\ve{x}} g=-\grad_{\ve\xi} g$.
Combining these with \eqref{g_L1_mu2} and
the bounds for $\norm{\lambda\,\pt_{\xi_2} g}{1\,;\Omega}$
and $\norm{\lambda\,\pt_{\xi_3} g}{1\,;\Omega}$ in \eqref{g_L1_mu3},
yields
\[
\norm{\lambda\, \pt_{x_1} g}{1\,;\Omega}
+\norm{\lambda\, D_{x_k} g}{1\,;\Omega}
+\norm{\lambda\, D_{\xi_k} g}{1\,;\Omega}
\leq C.
\]
Now, combining
$\pt_{x_1}\lambda=2q\eps^{-1}\lambda$ and
$\pt_q\lambda=2A\lambda\le 4\eps^{-1}\lambda$
with \eqref{g_L1_mu}, yields
\[
\norm{g\,\pt_{x_1}\lambda}{1\,;\Omega}
+\norm{g\,D_{x_k}\lambda}{1\,;\Omega}
+\norm{g\,D_{\xi_k}\lambda}{1\,;\Omega}
\leq C.
\]
Consequently, we get \eqref{D_g_lmbd}.
To estimate $\eps \widehat r\,D_{x_k}(\lambda\, \pt_{\xi_1} g)$,
note that it involves
$\eps \widehat r\,\pt_{x_k}(\lambda\, \pt_{\xi_1} g)
=-\eps \widehat r\,\lambda\, \pt^2_{\xi_1\xi_k} g$
for which we have a bound in \eqref{g_L1_mu3},
and also $\eps \widehat r\,\pt_q(\lambda\, \pt_{\xi_1} g)$,
for which we have a bound in \eqref{g_L1_mu4}. The desired bounds
for $\eps \widehat r\,D_{x_k}(\lambda\, \pt_{\xi_1} g)$
in \eqref{D_pt_lmbd} follow.
For $\eps \widehat r\,D_{\xi_k}\pt_{x_1}(\lambda g)$
in \eqref{D_pt_lmbd}, a calculation yields
$\eps \widehat r\,D_{\xi_k}\pt_{x_1}(\lambda g)
=\eps \widehat r\,D_{\xi_k}(\lambda\, \pt_{x_1} g)+2\widehat r\,D_{\xi_k}
( q \lambda\,g)$.
The first term is estimated similarly to
$\eps \widehat r\,D_{x_k}(\lambda\, \pt_{\xi_1} g)$ in \eqref{D_pt_lmbd}.
The remaining term $\widehat r\,D_{\xi_k}( q \lambda\,g)$ involves
$\widehat r\,\pt_{\xi_k}( q \lambda\,g)=q\,\widehat r\,\lambda\,\pt_{\xi_k} g$,
for which we have a bound in \eqref{g_L1_mu3}, and also
$\widehat r\,\pt_q( q \lambda\,g)=q \,\widehat r\,\pt_q( \lambda\,g)+\widehat r\,\lambda\,g$,
for which we have bounds in \eqref{g_L1_mu4} and \eqref{g_L1_mu}.
Consequently \eqref{D_pt_lmbd} is proved.
\end{proof}
\begin{lem}\label{lem_1_3}
Under the conditions of Lemma~\ref{lem:g_lmbd_bounds},
for some positive constant $c_1$ one has
\begin{gather}\label{lambda_g_1_3}
\|\lambda g(\ve{x};\cdot)\|_{2,1\,;[0\frac13]\times\R^2}
+\|D_{x_k}(\lambda g)(\ve{x};\cdot)\|_{1,1\,;[0\frac13]\times\R^2}
\le C\E^{-c_1\alpha/\eps},\quad k=2,\,3.
\end{gather}
\end{lem}
\begin{proof}
We imitate the proof of Lemma~\ref{lem:g_lmbd_bounds}, only now
$\xi_1<\frac13$ or $\widehat\xi_1< (\frac13-x_1)/\eps\le -\frac23/\eps$.
Thus instead of the sub-domains $\widehat\Omega_1'$ and $\widehat\Omega_2'$
we now consider $\widehat\Omega_1''$ and $\widehat\Omega_2''$
defined by
$\widehat\Omega_k'':=\widehat\Omega_k'\cap\{\widehat\xi_1< -(x_1-\frac13)/\eps\}$.
Thus in $\widehat\Omega_1''$ \eqref{eq:lambda_eqxi} remains valid with $q\ge\frac12\alpha$,
but now $\widehat r>\frac23/\eps$.
Therefore, when we integrate over $\widehat\Omega_1''$
(instead of $\widehat\Omega_1'$),
the integrals of type \eqref{star_lambda}, \eqref{star2_lambda}
become bounded by $C\E^{-c_1\alpha/\eps}$ for any fixed $ c_1<\frac1{8}$.
Next, when considering integrals over $\widehat\Omega_2''$
(instead of $\widehat\Omega_2'$),
note that $A-|\widehat\xi_1|\le-\frac23/\eps$ so
the quantity $\E^{2q(A-|\widehat\xi_1|)}$ in the definition \eqref{Q_def2} of $Q$
is now bounded by $\E^{-\frac23\alpha/\eps}$.
Consequently, the integrals of type \eqref{int_Omega2_main_lmbd}
over $\widehat\Omega_2''$ also become bounded by
$C\E^{-c_1\alpha/\eps}$.
\end{proof}
\begin{rem}\label{rem_q_var}
The estimates of Lemmas~\ref{lem:g0_bounds} and~\ref{lem:g_lmbd_bounds} remain valid
if we set $q:=\frac12a(\ve{x})$ or $q:=\frac12 a(\ve{\xi})$
in $g$, $\lambda$, and their derivatives (after the differentiation is performed).
\end{rem}
\section{Approximations $\bar G$ and $\tilde G$ for the Green's function~$G$}\label{sec:bounds_green_const}
We shall use two related cut-off functions $\omega_0$ and $\omega_1$
defined by
\begin{gather}\label{cut_off}
\omega_0(t) \in C^2(0,1),\quad\omega_0(t)=1\;\mbox{for }t\le{\ts\frac23},
\quad\omega_0(t)=0\;\mbox{for }t\ge{\ts\frac56};
\quad
\omega_1(t):=\omega_0(1-t),
\end{gather}
so $\omega_m(m)=1$, $\omega_m(1-m)=0$
and $\omega_m'(t)\bigr|_{t=0,1}=\omega_m''(t)\bigr|_{t=0,1}=0$ for $m=0,1$.
Our purpose in this section is to introduce and estimate
frozen-coefficient approximations
$\bar G$ and $\tilde G$ of $G$. We consider
the domain $\Omega=(0,1)\times\R^2$
in the first part of this section, and the domain $\Omega=(0,1)^3$ in the second part.
Note that although $\bar G$ and $\tilde G$ will be constructed as solution approximations
for the frozen-coefficient equations,
we shall see in Section~\ref{sec:main_proof} that they, in fact,
provide approximations to the Green's function $G$ for our original
variable-coefficient problem.
\subsection{Approximations
$\bar G$ and $\tilde G$
in the domain $\Omega=(0,1)\times\R^2$}\label{ssec:bounds_green_const_1}
To construct approximations
$\bar G$ and $\tilde G$, we employ the method of images
with an inclusion of the
cut-off functions of \eqref{cut_off}.
So, using the fundamental solution $g$ of \eqref{eq:def_g0}, we define
\begin{gather}\label{bar_tilde_G}
\bar G(\ve{x};\ve\xi):=\bar{\mathcal G}
\bigr|_{q=\frac12a(\ve{x})},
\qquad
\tilde G(\ve{x};\ve\xi):=\tilde {\mathcal G}
\bigr|_{q=\frac12a(\ve\xi)},
\end{gather}
\vspace{-0.8cm}
\begin{subequations}\label{bar_tilde_G20}
\begin{align}
\bar {\mathcal G}(\ve{x};\ve\xi;q)
&:=\frac{\E^{q\widehat\xi_{1,[x_1]}}}{4\pi\eps^2}
\left\{
\left[ \frac{\E^{-q\widehat r_{[ x_1]}}}{\widehat r_{[ x_1]}}
-\frac{\E^{-q\widehat r_{[ -x_1]}}}{\widehat r_{[ -x_1]}}\right]
-\left[ \frac{\E^{-q\widehat r_{[2-x_1]}}}{\widehat r_{[2-x_1]}}
-\frac{\E^{-q\widehat r_{[2+x_1]}}}{\widehat r_{[2+x_1]}}\right]\omega_1(\xi_1)
\right\},\label{bar_G_def}\\
\tilde {\mathcal G}(\ve{x};\ve\xi;q)
&:=\frac{\E^{q\widehat\xi_{1,[x_1]}}}{ 4\pi\eps^2}
\left\{
\left[ \frac{\E^{-q\widehat r_{[ x_1]}}}{\widehat r_{[ x_1]}}
-\frac{\E^{-q\widehat r_{[2-x_1]}}}{\widehat r_{[2-x_1]}}\right]
-\left[ \frac{\E^{-q\widehat r_{[ -x_1]}}}{\widehat r_{[ -x_1]}}
-\frac{\E^{-q\widehat r_{[2+x_1]}}}{\widehat r_{[2+x_1]}}\right]\omega_0(x_1)
\right\}\label{tilde_G_def}.
\end{align}
\end{subequations}
Note that $\bar G\bigr|_{\xi_1=0,1}=0$ and
$\tilde G\bigr|_{x_1=0,1}=0$
(the former observation
follows from $\widehat r_{[x_1]}=\widehat r_{[-x_1]}$ at $\xi_1=0$,
and $\widehat r_{[x_1]}=\widehat r_{[2-x_1]}$ and $\widehat r_{[-x_1]}=\widehat r_{[2+x_1]}$ at $\xi_1=1$).
We shall see shortly (see Lemma~\ref{lem_tilde_bar_w}) that
$\bar L^*_{\ve\xi}\bar G\approx L^*_{\ve\xi}G$
and $\tilde L_{\ve{x}}\tilde G\approx L_{\ve{x}} G$;
in this sense $\bar G$ and $\tilde G$ give approximations for $G$.
Rewrite the definitions of $\bar{\mathcal G}$ and $\tilde{\mathcal G}$
using the notation
\begin{subequations}
\begin{align}
g_{[d]
&:= g(d,x_2,x_3;\ve\xi;q)
={\frac{1}{ 4\pi\eps^2}}\, \frac{\E^{q(\widehat\xi_{1,[d]}-\widehat r_{[d]})}}{\widehat r_{[d]}},\label{g_x_brackets}\\
\lambda^{\pm}&:=\E^{2q(1\pm x_1)/{\eps}},
\qquad\quad p:=\E^{-2qx_1/{\eps}},\label{lmb_p_def}
\end{align}
\end{subequations}
and the observation that
\begin{gather}\label{q_d}
{\frac{1}{ 4\pi\eps^2}}\,\frac{\E^{q(\widehat\xi_{1,[x_1]}-\widehat r_{[d]})}}{\widehat r_{[d]}}
=\E^{q(d-x_1)/\eps}\,g_{[d]}\qquad\mbox{for}\;\; d=\pm x_1, 2\pm x_1.
\end{gather}
They yield
\begin{subequations}\label{bar_G_tilde_g}
\begin{align}
\bar{\mathcal G}(\ve{x};\ve\xi;q)
&= \left[g_{[x_1]}-p\, g_{[-x_1]}\right]
-\left[\lambda^- g_{[2-x_1]}-p\,\lambda^{\!+} g_{[2+x_1]}\right]\omega_1(\xi_1),\label{bar_G_g}\\
\tilde {\mathcal G}(\ve{x};\ve\xi;q)
&= \left[g_{[x_1]}-\lambda^- g_{[2-x_1]}\right]
-\left[p\, g_{[-x_1]}-p\,\lambda^{\!+} g_{[2+x_1]}\right]\omega_0(x_1).\label{tilde_G_g}
\end{align}
\end{subequations}
Note that $\lambda^\pm$ is obtained by replacing $x_1$ by $2\pm x_1$
in the definition \eqref{lambda_def} of $\lambda$.
In the next lemma, we estimate the functions
\begin{gather}\label{tilde_w_def}
\bar \phi(\ve{x};\ve\xi):=\bar L^*_{\ve\xi}\bar G-L^*_{\ve\xi}G,
\qquad
\tilde \phi(\ve{x};\ve\xi):=\tilde L_{\ve{x}}\tilde G-L_{\ve{x}}G.
\end{gather}
\begin{lem}\label{lem_tilde_bar_w}
Let $\ve{x}\in\Omega=(0,1)\times\R^2$. Then
for the functions $\bar\phi$ and $\tilde \phi$ of \eqref{tilde_w_def}, one has
\begin{gather}\label{tilde_bar_w}
\|\bar \phi(\ve{x};\cdot)\|_{1,1\,;\Omega}
+\|\pt_{x_2}\bar \phi(\ve{x};\cdot)\|_{1\,;\Omega}
+\|\pt_{x_3}\bar \phi(\ve{x};\cdot)\|_{1\,;\Omega}
+\|\tilde \phi(\ve{x};\cdot)\|_{1,1\,;\Omega}
\leq C\E^{-c_1\alpha/\eps}
\leq C.
\end{gather}
One also has
\begin{gather}\label{eq:bar_phi_boundary}
\bar\phi(\ve{x};\ve{\xi})|_{\ve{\xi}\in\pt\Omega}=0.
\end{gather}
\end{lem}
\begin{proof}
(i) First we prove the desired assertions for $\bar\phi$. By \eqref{bar_tilde_G},
throughout this part of the proof we set $q=\frac12a(\ve{x})\ge\frac12\alpha$.
Recall that $\bar g$ solves the differential equation \eqref{eq:Green_adj_const}
with the operator $\bar L^*_{\ve\xi}$.
Comparing the explicit formula for $\bar g$ in \eqref{bar_tilde_g_def}
with the notation \eqref{g_x_brackets} implies that
$\bar L^*_{\ve\xi}g_{[d]}=\delta(\xi_1-d)\,\delta(\xi_2-x_2)\,\delta(\xi_3-x_3)$.
So, by \eqref{eq:Green_adj}, $\bar L^*_{\ve\xi}g_{[x_1]}=L_{\ve\xi}^*G$,
and also $\bar L^*_{\ve\xi}g_{[d]}=0$ for $d=-x_1,2\pm x_1$ and all $\ve\xi\in\Omega$.
Now, by \eqref{bar_G_g}, we conclude that
$\bar\phi=-\bar L^*_{\ve\xi}[\omega_1(\xi_1) \bar {\mathcal G}_2]$
where $\bar {\mathcal G}_2:=\lambda^- g_{[2-x_1]}-p\,\lambda^{\!+} g_{[x_1+2]}$,
and $\bar L^*_{\ve\xi}\bar {\mathcal G}_2=0$ for $\ve\xi\in\Omega$.
From these observations,
$\bar\phi=2\eps\omega_1'(\xi_1)\pt_{\xi_1} \bar {\mathcal G}_2
+[\eps\omega_1''(\xi_1)-2q\omega_1'(\xi_1)]\bar {\mathcal G}_2$.
The definition \eqref{cut_off} of $\omega_1$ implies that
$\bar\phi$ vanishes at $\xi_1=0$ and for $\xi_1\ge\frac13$.
This implies the desired assertion \eqref{eq:bar_phi_boundary}.
Furthermore, we now get
\begin{multline*}
\norm{\bar \phi(\ve{x};\cdot)}{1,1\,;\Omega}
+\norm{\pt_{x_2}\bar \phi(\ve{x};\cdot)}{1\,;\Omega}
+\norm{\pt_{x_3}\bar \phi(\ve{x};\cdot)}{1\,;\Omega}\\
\leq C \bigl(
\norm{\bar {\mathcal G}_2(\ve{x};\cdot)}{2,1\,;[0\frac13]\times\R}
+\norm{D_{x_2}\bar {\mathcal G}_2(\ve{x};\cdot)}{1,1\,;[0\frac13]\times\R}
+\norm{D_{x_3}\bar {\mathcal G}_2(\ve{x};\cdot)}{1,1\,;[0\frac13]\times\R}
\bigr).
\end{multline*}
Combining this with the bounds \eqref{lambda_g_1_3} for the terms $\lambda^\pm g_{[2\pm x_1]}$
of $\bar {\mathcal G}_2$, and the observation that
$|D_{x_2}p|+|D_{x_3}p|\le C|\pt_q p|\le C$ and $\pt_{\xi_k} p=0,\,k=1,2,3$, yields
our assertions for $\bar\phi$ in \eqref{tilde_bar_w}.
(ii) Now we prove the desired estimate \eqref{tilde_bar_w} for $\tilde \phi$.
By \eqref{bar_tilde_G}, throughout this part of the proof we set $q=\frac12a(\ve\xi)\ge\frac12\alpha$.
Comparing the notation \eqref{g_x_brackets} with
the explicit formula for $\tilde g$ in \eqref{bar_tilde_g_def}, we rewrite
\eqref{eq:Green_prim_const}
as $\tilde L_{\ve{x}}g_{[x_1]}=\delta(\ve{x}-\ve\xi)$.
So $\tilde L_{\ve{x}}g_{[x_1]}=L_{\ve{x}}G$, by~\eqref{eq:Green_prim}.
Next, for each value $d=-x_1,2\pm x_1$ respectively set
$s=-\xi_1,\mp(2-\xi_1)$. Now by \eqref{eq:def_g0}, one has
$\widehat r_{[d]}=\sqrt{(s-x_1)^2+(\xi_2-x_2)^2+(\xi_3-x_3)^2}/\eps$
so $g(\ve{x};s,\xi_2,\xi_3;q)=\ds\frac{1}{4\pi\eps^2}\frac{\E^{q(s-x_1)/\eps-q\widehat r_{[d]}}}{\widehat r_{[d]}}$.
Note that $\tilde L_{\ve{x}}g(\ve{x};s,\xi_2,\xi_3;q)=\delta(x_1-s)\,\delta(x_2-\xi_2)\,\delta(x_3-\xi_3)$
and none of our three values of $s$ is in $[0,1]$ (i.e. $\delta(s-x_1)=0$).
Consequently,
$\tilde L_{\ve{x}}
\Bigl[\ds\frac{\E^{q(\widehat\xi_{1,[x_1]}-\widehat r_{[d]})}}
{\widehat r_{[d]}}\Bigr]=0$
for all $\ve{x}\in\Omega$.
Comparing \eqref{tilde_G_def} and \eqref{tilde_G_g}, we now conclude that
$\tilde\phi=-\tilde L_{\ve{x}}[\omega_0(\xi_1) \tilde {\mathcal G}_2]$
where $\tilde{\mathcal G}_2:=p\, g_{[-x_1]}-p\,\lambda^{\!+} g_{[2+x_1]}$
and $\tilde L_{\ve{x}}\tilde{\mathcal G}_2=0$ for $\ve{x}\in\Omega$.
From these observations,
$\tilde\phi =2\eps\omega_0'(x_1)\,\pt_{x_1} \tilde{\mathcal G}_2
+[\eps\omega_0''(x_1)+2q\omega_0'(x_1)]\tilde{\mathcal G}_2$.
As the definition~\eqref{cut_off} of $\omega_0$ implies that $\tilde\phi$
vanishes for $x_1\le\frac23$, we have
\[
\norm{\tilde\phi(\ve{x};\cdot)}{1,1\,;\Omega}
\leq C \max_{\stackrel{\ve{x}\in[\frac23,1]\times\R^2}
{k\,=\,0,1}}
\norm{\pt^k_{x_1}\tilde {\mathcal G}_2(\ve{x};\cdot)}{1,1\,;\Omega}\,.
\]
Here $\tilde G_2$ is smooth and has no singularities for $x_1\in[\frac23,1]$
(because $\widehat r_{[2+x_1]}\geq \widehat r_{[-x_1]}\geq \frac23\eps^{-1}$ for $x\in[\frac23,1]$).
Note that $\norm{\pt^k_{x_1} g_{[-x_1]}}{1,1\,;\Omega}\le C \eps^{-2}$,
and $\norm{\pt^k_{x_1} (\lambda^{\!+}g_{[2+x_1]})}{1,1\,;\Omega}\le C \eps^{-2}$
(these two estimates are similar to the ones in
Lemmas~\ref{lem:g0_bounds} and~\ref{lem:g_lmbd_bounds},
but easier to deduce as they are not sharp).
We combine these two bounds with
$|\pt^k_{x_1} \pt^l_{\xi_1}\pt^m_{\xi_2}\pt^n_{\xi_3} p|
\leq C \eps^{-2}p
= C \eps^{-2}\E^{-2qx_1/{\eps}}$ for $k,\, l+m+n\leq 1$.
As for $x_1\ge\frac23$ we enjoy the bound
$\E^{-2qx_1/{\eps}}
\leq \E^{-\frac23\alpha/{\eps}}
\leq C\eps^4 \E^{-\frac12\alpha/{\eps}}$, the desired estimate for $\tilde\phi$ follows.
\end{proof}
\begin{lem}\label{lem:tilde_bar_G}
Let the function $R=R(\ve{x};\ve\xi)$ be such that
$|R|\le C\min\{\eps\widehat r_{[\ve{x}]},1\}$.
The functions $\bar G$ and $\tilde G$ of \eqref{bar_tilde_G}, \eqref{bar_G_tilde_g}
satisfy
\begin{subequations}
\begin{align}
\norm{\bar G(\ve{x};\cdot)}{1\,;\Omega}+\norm{\tilde G(\ve{x};\cdot)}{1\,;\Omega}
&\le C,\label{G_bar_t_1}\\
\norm{\pt_{\xi_1}\bar G(\ve{x};\cdot)}{1\,;\Omega}
&\le C(1+|\ln\eps|),\label{bar_G_xi}\\
\norm{\pt_{\xi_k}\bar G(\ve{x};\cdot)}{1\,;\Omega}
&\le C\eps^{-1/2},\quad k=2,\,3,\label{bar_G_eta}\\
\norm{(R\,\pt_{\xi_1}\bar G)(\ve{x};\cdot)}{1\,;\Omega}
+\eps^{1/2}\norm{(R\,\pt^2_{\xi_1\xi_k}\bar G)(\ve{x};\cdot)}{1\,;\Omega}
&\le C,\quad\,\qquad k=2,\,3,\label{bar_G_xi_eta}
\intertext{and for any ball $B(\ve{x}';\rho)$ of radius $\rho$ centered at any
$\ve{x}'\in[0,1]\times\R^2$, one has}
|\bar G(\ve{x};\cdot)|_{1,1\,;B(\ve{x}';\rho)\cap\Omega}
&\le C\eps^{-1}\rho,\label{bar_G_1_1_B}
\end{align}
while for the ball $B(\ve{x};\rho)$ of radius $\rho$ centered at $\ve{x}$,
we have
\begin{align}
\norm{\pt^2_{\xi_1}\bar G(\ve{x};\cdot)}{1\,;\Omega\setminus B(\ve{x};\rho)}
&\le C\eps^{-1}\ln(2+\eps/\rho),\label{bar_G_xi_xi_2}\\
\norm{\pt^2_{\xi_k}\bar G(\ve{x};\cdot)}{1\,;\Omega\setminus B(\ve{x};\rho)}
&\le C\eps^{-1}(\ln(2+\eps/\rho)+|\ln\eps|),\quad k=2,\,3.\label{bar_G_eta_eta_2}
\end{align}
Furthermore, we have for $k=2,\,3$
\begin{align}
\norm{\pt_{x_k}\bar G(\ve{x};\cdot)}{1\,;\Omega}
+\norm{(R\,\pt^2_{\xi_1 x_k}\bar G)(\ve{x};\cdot)}{1\,;\Omega}
&\le C\eps^{-1/2},\label{bar_G_y}\\
\norm{\pt_{\xi_k}\tilde G(\ve{x};\cdot)}{1\,;\Omega}
&\le C\eps^{-1/2},\label{G_tilde_eta}\\
\int_0^1\!\bigl(\norm{(R\,\pt^2_{x_1\xi_k}\tilde G)(\ve{x};\cdot)}{1\,;\Omega}+
\norm{\pt_{x_1}\tilde G(\ve{x};\cdot)}{1\,;\Omega}\bigr)\,dx_1
&\le C\eps^{-1/2}.\label{G_tilde_eta_x}
\end{align}
\end{subequations}
\end{lem}
\begin{proof}
Throughout the proof,
whenever $k$ appears in any relation,
it will be understood to be valid for $k=2,3$.
First, note that $\widehat r_{[-x_1]}\geq \widehat r_{[x_1]}$
and $\widehat r_{[2\pm x_1]}\ge \widehat r_{[x_1]}$ for all $\ve\xi\in\Omega$, therefore
\begin{gather}\label{R_ext}
|R|\leq C\,\min\bigl\{\eps\widehat r_{[x_1]},\,\eps\widehat r_{[-x_1]},\,\eps\widehat r_{[2-x_1]},\,\eps\widehat r_{[2+x_1]},
\,1\bigr\}.
\end{gather}
Note also that in view of Remark~\ref{rem_q_var},
all bounds of Lemma~\ref{lem:g0_bounds} apply to
the components $g_{[\pm x_1]}$
and all bounds of Lemma~\ref{lem:g_lmbd_bounds} apply to
the components $\lambda^\pm g_{[2\pm x_1]}$
of $\bar{\mathcal G}$ and $\tilde{\mathcal G}$ in \eqref{bar_G_tilde_g}.
\underline{\it Asterisk notation.}
In some parts of this proof, when discussing derivatives of $\bar {\mathcal G}$,
we shall use the notation $\bar {\mathcal G}^\ast$
prefixed by some differential operator, e.g., $\pt_{x_1} \bar {\mathcal G}^\ast$.
This will mean that the differential operator is applied only to the terms of the
type $g_{[d\pm x_1]}$, e.g., $\pt_{x_1} \bar {\mathcal G}^\ast$ is obtained by replacing
each of the four terms $g_{[d\pm x_1]}$ in the definition \eqref{bar_G_g}
of $\bar {\mathcal G}$ by $\pt_{x_1} g_{[d\pm x_1]}$ respectively.
\begin{enumerate}
\item The first desired estimate \eqref{G_bar_t_1}
follows from the bound \eqref{g_L1} for $g_{[\pm x_1]}$
and the bound \eqref{g_L1_mu} for $\lambda^\pm g_{[2\pm x_1]}$
combined with $|p|\le 1$ and $|\omega_{0,1}|\le 1$
(in fact, the bound for $\bar G$ can obtained
by imitating the proof of Lemma~\ref{lem_G_L1}).
\item Rewrite \eqref{bar_G_g} as
\[
\bar {\mathcal G}=\bar {\mathcal G}_1-\omega_1(\xi_1)\bar {\mathcal G}_2,
\quad\mbox{where}\quad
\bar {\mathcal G}_1:=g_{[x_1]}-p\, g_{[-x_1]},
\quad
\bar {\mathcal G}_2:=\lambda^- g_{[2-x_1]}-p\,\lambda^{\!+} g_{[2+x_1]}.
\]
As $q=\frac12a(\ve{x})$ in $\bar G$
(i.e. $p$ and $\lambda^\pm$ in $\bar G$ do not involve $\ve\xi$),
one gets
\[
\pt_{\xi_1} \bar G=\pt_{\xi_1}\bar {\mathcal G}^\ast-\omega_1'(\xi_1)\bar {\mathcal G}_2,
\qquad
\pt_{\xi_k} \bar G=\pt_{\xi_k}\bar {\mathcal G}^\ast,
\qquad
\pt^2_{\xi_1\xi_k} \bar G=\pt^2_{\xi_1\xi_k}\bar {\mathcal G}^\ast
-\omega_1'(\xi_1)\pt_{\xi_k}\bar {\mathcal G}^\ast_2.
\]
Now the desired estimate \eqref{bar_G_xi}
follows from the bound \eqref{g_xi_L1} for $\pt_{\xi_1} g_{[\pm x_1]}$,
the bound \eqref{g_L1_mu2} for $\lambda^\pm \,\pt_{\xi_1} g_{[2\pm x_1]}$,
and the bound \eqref{g_L1_mu} for $\lambda^\pm g_{[2\pm x_1]}$.
Similarly, our next assertion \eqref{bar_G_eta}
follows from the bound \eqref{g_eta_L1} for $\pt_{\xi_k} g_{[\pm x_1]}$,
and the bound \eqref{g_L1_mu3} for $\lambda^\pm \pt_{\xi_k}g_{[2\pm x_1]}$.
The next estimate \eqref{bar_G_xi_eta} is deduced using
\[
|R\,\pt_{\xi_1} \bar G|
\leq |R \,\pt_{\xi_1}\bar {\mathcal G}_1^\ast|
+C|\pt_{\xi_1}\bar {\mathcal G}_2^\ast|
+C|\bar {\mathcal G}_2|,
\quad
|R\,\pt^2_{\xi_1\xi_k} \bar G|
\leq |R\,\pt^2_{\xi_1\xi_k}\bar {\mathcal G}^\ast|
+C|\pt_{\xi_k}\bar {\mathcal G}^\ast_2|.
\]
Here, in view of \eqref{R_ext}, the term $R \,\pt_{\xi_1}\bar {\mathcal G}_1^\ast$
is estimated using the bound \eqref{g_xi_R_L1} for
$\eps\widehat r_{[\pm x_1]} \pt_{\xi_1} g_{[\pm x_1]}$,
while the terms $R\,\pt^2_{\xi_1\xi_k}\bar {\mathcal G}^\ast$
are estimated using the bound \eqref{g_eta_x_L1} for
$\eps\widehat r_{[\pm x_1]}\pt^2_{\xi_1\xi_k}g_{[\pm x_1]}$
and the bound \eqref{g_L1_mu3} for
$\lambda^\pm \eps\widehat r_{[2\pm x_1]}\pt^2_{\xi_1\xi_k}g_{[2\pm x_1]}$.
The remaining terms $\pt_{\xi_1}\bar {\mathcal G}_2^\ast$,
$\bar {\mathcal G}_2$
and $\pt_{\xi_k}\bar{\mathcal G}^\ast_2$
appear in $\pt_{\xi_1}\bar G$ and $\pt_{\xi_k}\bar G$,
so have been bounded when obtaining \eqref{bar_G_xi}, \eqref{bar_G_eta}.
\item The next assertion \eqref{bar_G_1_1_B} is proved similarly to
\eqref{bar_G_xi} and \eqref{bar_G_eta}, only using
the bound \eqref{g_eta_ball} for $g_{[\pm x_1]}$
and the bound \eqref{g_L1_mu5} for $\lambda^\pm g_{[2\pm x_1]}$.
\item As $q=\frac12a(\ve{x})$ in $\bar G$, then
$\pt^2_{\xi_m} \bar G=\pt^2_{\xi_m}\bar {\mathcal G}^\ast$, $m=1,\,2,\,3$,
and the assertions \eqref{bar_G_xi_xi_2} and \eqref{bar_G_eta_eta_2}
immediately follow from the bounds \eqref{g_xi2_L1} and \eqref{g_eta2_L1} for
$\pt^2_{\xi_m}g_{[\pm x_1]}$ combined with
the bounds \eqref{g_L1_mu6} for $\lambda^\pm \pt^2_{\xi_m}g_{[2\pm x_1]}$
where $m=1,\,2,\,3$.
\item As $q=\frac12a(\ve{x})$ in $\bar G$, so
using the operator $D_{x_k}$ of \eqref{D_ops}, one gets
\begin{align*}
\pt_{x_k} \bar G
=D_{x_k}\bigl[g_{[x_1]}-p \,g_{[-x_1]}\bigr]^\ast
-\omega_1(\xi_1)\,
\bigl[D_{x_k}(\lambda^- g_{[2-x_1]})-p\, D_{x_k}(\lambda^{\!+} g_{[2+x_1]})\bigr]\\
-{\ts\frac12} \pt_{x_k} a(\ve{x})\cdot \pt_q p\cdot
\bigl[ g_{[-x_1]}-\omega_1(\xi_1)\lambda^{\!+} g_{[2+x_1]}\bigr],
\end{align*}
where $|\pt_q p|\le C$ by \eqref{lmb_p_def}
(and we used the previously defined asterisk notation).
Now, $\pt_{x_k} \bar G$ is estimated using the
bound \eqref{D_g} for $D_{x_k}g_{[\pm x_1]}$ and
the bound \eqref{D_g_lmbd} for $D_{x_k}(\lambda^\pm g_{[2\pm x_1]})$.
For the term $g_{[-x_1]}$ in $\pt_{x_k} \bar G$
we use the bound \eqref{g_L1}, and for the term $\lambda^{\!+} g_{[2+x_1]}$
the bound \eqref{g_L1_mu}.
Consequently, one gets the desired bound \eqref{bar_G_y}
for $D_{x_k}\bar G^\ast$.
To estimate $R\,\pt^2_{\xi_1 x_k} \bar G$, $k=2,\,3$, a calculation shows that
\begin{align*}
\pt^2_{\xi_1 x_k} \bar G
=(D_{x_k}\pt_{\xi_1}) \bigl[g_{[x_1]}-p \,g_{[-x_1]}\bigr]^\ast
\!\!-\omega_1(\xi_1)\,
\bigl[D_{x_k}(\lambda^- \pt_{\xi_1} g_{[2-x_1]})-p\, D_{x_k}(\lambda^{\!+}\pt_{\xi_1} g_{[2+x_1]})\bigr]\\
-{\ts\frac12} \pt_{x_k} a(\ve{x})\cdot \pt_q p\cdot
\bigl[\pt_{\xi_1} g_{[-x_1]}
-\omega_1(\xi_1)\lambda^{\!+}\pt_{\xi_1} g_{[2+x_1]}\bigr]
-\omega_1'(\xi_1)\,\pt_{x_k}\bar G_2,
\end{align*}
where $\bar G_2:=\bar{\mathcal G}_2\bigr|_{q=a(\ve{x})/2}$.
The assertion \eqref{bar_G_y} for $R\,\pt^2_{\xi_1 x_k} \bar G$
is now deduced as follows. In view of \eqref{R_ext},
we employ the bound \eqref{D_pt_g} for the terms
$\eps\widehat r_{[\pm x_1]}D_{x_k}\pt_{\xi_1} g_{[\pm x_1]}$
and the bound \eqref{D_pt_lmbd} for the terms
$\eps \widehat r_{[2\pm x_1]}\,D_{x_k}(\lambda^\pm\, \pt_{\xi_1} g_{[2\pm x_1]})$.
For the remaining terms (that appear in the second line)
we use $|R|\le C$ and $|\pt_q p|\le C$.
Then we combine the bound \eqref{g_xi_L1} for $\pt_{\xi_1} g_{[-x_1]}$
and the bound \eqref{g_L1_mu2} for $\lambda^{\!+}\pt_\xi g_{[2+x_1]}$.
The term $\pt_{x_k}\bar G_2$ is a part of $\pt_{x_k}\bar G$, which was estimated above,
so for $\pt_{x_k}\bar G_2$
we have the same bound as for $\pt_{x_k}\bar G$ in \eqref{bar_G_y}.
This observation completes the proof of the bound
for $R\,\pt^2_{\xi_1 x_k} \bar G$ in \eqref{bar_G_y}.
\item We now proceed to estimating derivatives of $\tilde G$,
so $q=\frac12a(\ve\xi)$ in this part of the proof.
Let $\tilde {\mathcal G}^\pm:=g_{[\pm x_1]}-\lambda^{\mp} g_{[2\mp x_1]}$.
Then \eqref{tilde_G_g}, \eqref{lmb_p_def} imply that
$\tilde {\mathcal G}=\tilde {\mathcal G}^+-p_0\tilde {\mathcal G}^-$, where
$p_0:=\omega_0(x_1)\,p=\omega_0(x_1)\,\E^{-2qx_1/\eps}$.
Note that
\[
D_{\xi_k} p_0
={\ts\frac12}\pt_{\xi_k} a(\ve\xi)\cdot
(-2x_1/\eps)\,p_0
\qquad
\pt_{x_1} p_0=[\omega_0'(x_1)-(2q/\eps)\,\omega_0(x_1)]\,\E^{-2qx_1/{\eps}}.
\]
Combining this with $|(-2x_1/\eps)\,p_0|\le C \E^{-qx_1/{\eps}}$
and $q\ge \frac12 \alpha$ yields
\begin{gather}\label{p_0}
|D_{\xi_k} p_0|\le C,\qquad
\int_0^1\!\bigl(|\pt_{x_1} p_0|+|D_{\xi_k} \pt_{x_1} p_0|\bigr)\,dx_1
\leq \int_0^1\!\bigl(C\eps^{-1}\E^{-\frac12\alpha x_1/{\eps}}\bigr)\,dx_1
\leq C.
\end{gather}
Furthermore, we claim that
\begin{gather}\label{tilde_cal_G}
\norm{\tilde {\mathcal G}^-}{1\,;\Omega}\le C,
\qquad
\norm{\pt_{x_1}\tilde {\mathcal G}^\pm}{1\,;\Omega}\le C(1+|\ln\eps|),
\qquad
\norm{D_{\xi_k}\tilde {\mathcal G}^\pm}{1\,;\Omega}\le C\eps^{-1/2}.
\end{gather}
Here the first estimate
follows from the bounds \eqref{g_L1}, \eqref{g_L1_mu} for
the terms $g_{[-x_1]}$ and $\lambda^{\!+}g_{[2+x_1]}$.
The estimate for $\pt_{x_1}\tilde {\mathcal G}^\pm$ in \eqref{tilde_cal_G} follows from
the bound \eqref{g_x_L1} for $\pt_{x_1} g_{[\pm x_1]}$ and
the bound \eqref{D_g_lmbd} for $\pt_{x_1}(\lambda^{\pm}g_{[2\pm x_1]})$.
Similarly, the estimate for $D_{\xi_k}\tilde {\mathcal G}^\pm$
in \eqref{tilde_cal_G} is obtained using
the bound \eqref{D_g} for $D_{\xi_k} g_{[\pm x_1]}$ and
the bound \eqref{D_g_lmbd} for $D_{\xi_k}(\lambda^{\pm}g_{[2\pm x_1]})$.
Next, a calculation shows that
\[
\pt_{\xi_k}\tilde G
= D_{\xi_k}\tilde {\mathcal G}^+
-p_0\,D_{\xi_k}\tilde {\mathcal G}^-
-D_{\xi_k} p_0\cdot\tilde {\mathcal G}^-,
\qquad
\pt_{x_1}\tilde G
= \pt_{x_1}\tilde {\mathcal G}^+
-p_0\,\pt_{x_1}\tilde {\mathcal G}^-
-\pt_{x_1} p_0\cdot\tilde {\mathcal G}^-.
\]
Combining these with \eqref{p_0}, \eqref{tilde_cal_G} yields
\eqref{G_tilde_eta} and the bound for $\pt_{x_1}\tilde G$ in \eqref{G_tilde_eta_x}.
To establish the estimate for $R\,\pt^2_{x_1\xi_k}\tilde G$
in \eqref{G_tilde_eta_x},
note that
\[
\pt^2_{x_1\xi_k}\tilde G
= D_{\xi_k}\pt_{x_1} \tilde {\mathcal G}^+
-p_0\cdot D_{\xi_k}\pt_{x_1} \tilde {\mathcal G}^-\!
-\pt_{x_1} p_0\cdot D_{\xi_k} \tilde {\mathcal G}^-\!
-\pt_{\xi_k} p_0\cdot \pt_{x_1} \tilde {\mathcal G}^-\!
-D_{\xi_k}\pt_{x_1} p_0\cdot \tilde {\mathcal G}^-\!.
\]
In view of \eqref{R_ext}, \eqref{p_0} and \eqref{tilde_cal_G},
it now suffices to show that
$\|R\,D_{\xi_k}\pt_{x_1} \tilde {\mathcal G}^\pm\|_{1\,;\Omega}\le C\eps^{-1/2}$.
This latter estimate immediately
follows from the bound \eqref{D_pt_g}
for the terms $\eps\widehat r_{[\pm x_1]}\, D_{\xi_k}\pt_{x_1} g_{[\pm x_1]}$
and the bound \eqref{D_pt_lmbd} for the terms
$\eps\widehat r_{[\pm x_1]}\, D_{\xi_k}\pt_{x_1} (\lambda^\pm g_{[2\pm x_1]})$.
This completes the proof of \eqref{G_tilde_eta_x}.
\end{enumerate}
\end{proof}
\subsection{Approximations for the Green's function $G$ in the domain $\Omega=(0,1)^3$}\label{ssec:bounds_green_const_2}
We now define approximations, denoted by $\bar G_\cube$ and $\tilde G_\cube$,
for the Green's function $G$ in
our original domain $\Omega=(0,1)^3$. For this, we use
the approximations $\bar G$ and $\tilde G$ of \eqref{bar_tilde_G}, \eqref{bar_tilde_G20}
for the domain $(0,1)\times\R^2$
and again employ the method of images with an inclusion of the cut-off
functions of \eqref{cut_off} in a two-step process as follows:
\begin{subequations}\label{eq:def_tilde_bar_G}
\begin{align}
\!\bar G_{\mbox{\tiny$\Box$}}(\ve{x};\ve\xi)
&:= \bar G(\ve{x};\ve\xi)
\hspace{-0.0cm}&&-\omega_0(\xi_2)\,\bar G(\ve{x};\xi_1,-\xi_2,\xi_3)
\hspace{0.1cm}&&-\omega_1(\xi_2)\,\bar G(\ve{x};\xi_1,2-\xi_2,\xi_3),\notag\\
\!\bar G_\cube(\ve{x};\ve\xi)
&:= \bar G_{\mbox{\tiny$\Box$}}(\ve{x};\ve\xi)
\hspace{-1cm}&&-\omega_0(\xi_3)\,\bar G_{\mbox{\tiny$\Box$}}(\ve{x};\xi_1,\xi_2,-\xi_3)
\hspace{-1cm}&&-\omega_1(\xi_3)\,\bar G_{\mbox{\tiny$\Box$}}(\ve{x};\xi_1,\xi_2,2-\xi_3),\\
\!\tilde G_{\mbox{\tiny$\Box$}}(\ve{x};\ve\xi)
&:= \tilde G(\ve{x};\ve\xi)
\hspace{-1cm}&&-\omega_0(x_2)\,\tilde G(x_1,-x_2,x_3;\ve\xi)
\hspace{-1cm}&&-\omega_1(x_2)\,\tilde G(x_1,2-x_2,x_3;\ve\xi),\notag\\
\!\tilde G_\cube(\ve{x};\ve\xi)
&:= \tilde G_{\mbox{\tiny$\Box$}}(\ve{x};\ve\xi)
\hspace{-1cm}&&-\omega_0(x_3)\,\tilde G_{\mbox{\tiny$\Box$}}(x_1,x_2,-x_3;\ve\xi)
\hspace{-1cm}&&-\omega_1(x_3)\,\tilde G_{\mbox{\tiny$\Box$}}(x_1,x_2,2-x_3;\ve\xi).
\end{align}
\end{subequations}
Then $\bar G_\cube\bigr|_{\xi_1=0,1}=0$
and $\tilde G_\cube\bigr|_{x_1=0,1}=0$
(as this is valid for $\bar G$ and $\tilde G$, respectively),
and furthermore, by \eqref{cut_off}, we have
$\bar G_\cube\bigr|_{\xi_k=0,1}=0$
and $\tilde G_\cube\bigr|_{x_k=0,1}=0$
for $k=2,3$.
\begin{rem}\label{rem:strip_to_square}
Lemmas~\ref{lem_tilde_bar_w} and \ref{lem:tilde_bar_G}
of the previous section remain valid if $\Omega$ is understood as $(0,1)^3$, and
$\bar G$ and $\tilde G$ are replaced by $\bar G_\cube$ and $\tilde G_\cube$,
respectively, in the definition \eqref{tilde_w_def} of $\bar\phi$ and $\tilde\phi$
and in the lemma statements.
This is shown by imitating the proofs of these two lemmas.
We leave out the details and only note that the application of the method of images
in the $\xi_2$- and $\xi_3$- ($x_2$- and $x_3$-) directions
is relatively straightforward as an inspection of \eqref{eq:def_g0}
shows that in these directions, the fundamental solution $g$
is symmetric and exponentially decaying away from the singular point.
\end{rem}
As $\bar G_\cube$ and $\tilde G_\cube$ in the domain $\Omega=(0,1)^3$
enjoy the same properties as $\bar G$ and $\tilde G$ in the domain
$(0,1)\times\R^2$, we shall sometimes skip the subscript
\cube{} when there is no ambiguity.
\section{Proof of Theorem~\ref{thm:main} for $\Omega=(0,1)^3$\\
(general variable-coefficient case)}\label{sec:main_proof}
We are now ready to establish our main result, Theorem~\ref{thm:main},
for the original variable-coefficient problem \eqref{eq:Lu} in the domain
$\Omega=(0,1)^3$.
In Section~\ref{sec:bounds_green_const}, we have already
obtained various bounds for the approximations $\tilde G_\cube$
and $\bar G_\cube$ of $G$ in $\Omega=(0,1)^3$.
So now we consider the two functions
\[
\tilde v(\ve{x};\ve\xi):=[G-\tilde G_\cube](\ve{x};\ve\xi),
\qquad
\bar v(\ve{x};\ve\xi)=[G-\bar G_\cube](\ve{x};\ve\xi).
\]
Throughout this section, we shall skip the subscript \cube{}
as we always deal with the domain $\Omega=(0,1)^3$.
Note that, by \eqref{tilde_w_def}, we have
$L_{\ve{x}}\tilde v=L_{\ve{x}}[G-\tilde G]
= [\tilde L_{\ve{x}}-L_{\ve{x}}]\tilde G-\tilde\phi$,
and similarly
$L^*_{\ve\xi}\bar v=L^*_{\ve\xi}[G-\bar G]
= [\bar L^*_{\ve\xi}-L^*_{\ve\xi}]\bar G-\bar\phi$.
Consequently, the functions $\tilde v$ and $\bar v$ are solutions of the following problems:
\begin{subequations}
\begin{align}
L_{\ve{x}} \tilde v(\ve{x};\ve\xi)
&= \tilde h(\ve{x};\ve\xi)\;\;\mbox{for}\;\ve{x}\in\Omega,
\qquad \tilde v(\ve{x};\ve\xi)=0\;\;\mbox{for}\; \ve{x}\in\pt\Omega,\label{prob_tilde_v}\\
L^*_{\ve\xi} \bar v(\ve{x};\ve\xi)
&= \bar h(\ve{x};\ve\xi)\;\;\mbox{for}\;\ve\xi\in\Omega,
\qquad \bar v(\ve{x};\ve\xi)=0\;\;\mbox{for}\; \ve\xi\in\pt\Omega.\label{prob_bar_v}
\end{align}
\end{subequations}
Here the right-hand sides are given by
\begin{subequations}
\begin{align}
\tilde h(\ve{x};\ve\xi)
&:= \pt_{x_1}\{R\,\tilde G\}(\ve{x};\ve\xi)
-b(\ve{x})\,\tilde G(\ve{x};\ve\xi)
-\tilde \phi(\ve{x};\ve\xi),\label{tilde_bar_h}\\
\bar h(\ve{x};\ve\xi)
&:= \{R\,\pt_{\xi_1}\bar G\}(\ve{x};\ve\xi)
-b(\ve\xi)\,\bar G(\ve{x};\ve\xi)
-\bar \phi(\ve{x};\ve\xi),\label{bar_h}
\end{align}
\end{subequations}
where
\begin{gather}\label{R_def}
R(\ve{x};\ve\xi):=a(\ve{x})-a(\ve\xi),\qquad\mbox{so}\;\;
|R|\le C\min\{\eps\widehat r_{[x_1]},1\}.
\end{gather}
Applying the solution representation formulas \eqref{eq:sol_prim} and
\eqref{eq:sol_adj} to problems \eqref{prob_tilde_v} and \eqref{prob_bar_v},
respectively, one gets
\begin{subequations}
\begin{align}
\tilde v(\ve{x};\ve\xi)
&= \iiint_\Omega G(\ve{x};\ve{s})\,\tilde h(\ve{s};\ve\xi)\,d\ve{s},\label{tilde_v}\\
\bar v(\ve{x};\ve\xi)
&= \iiint_\Omega G(\ve{s};\ve\xi)\,\bar h(\ve{x};\ve{s})\,d\ve{s}.\label{bar_v}
\end{align}
\end{subequations}
We now proceed to the {proof of Theorem~\ref{thm:main}}.
\smallskip
\begin{proof}
Throughout the proof,
whenever $k$ appears in any relation,
it will be understood to be valid for $k=2,3$.
(i)
First we establish \eqref{eq:thm:G_eta}.
Note that, the bounds \eqref{G_tilde_eta} and \eqref{bar_G_y}
for $\pt_{\xi_k}\tilde G$ and $\pt_{x_k}\bar G$,
respectively, it suffices to show that
$ \norm{\pt_{\xi_k}\tilde v(\ve{x};\cdot)}{1\,;\Omega}
+\norm{\pt_{x_k}\bar v(\ve{x};\cdot)}{1\,;\Omega}\le C\eps^{-1/2}$.
Applying $\pt_{\xi_k}$ to \eqref{tilde_v} and $\pt_{x_k}$ to \eqref{bar_v}, we arrive at
\begin{align*}
\pt_{\xi_k}\tilde v(\ve{x};\ve\xi)
&= \iiint_\Omega \!G(\ve{x};\ve{s})\,\pt_{\xi_k}\tilde h(\ve{s};\ve\xi)\,d\ve{s},\\
\pt_{x_k}\bar v(\ve{x};\ve\xi)
&=\iiint_\Omega\!G(\ve{s};\ve\xi)\,\pt_{x_k}\bar h(\ve{x};\ve{s})\,d\ve{s}.
\end{align*}
From this, a calculation shows that
\begin{align*}
\norm{\pt_{\xi_k}\tilde v(\ve{x};\cdot)}{1\,;\Omega}
&\le \Bigl(\sup_{s_1\in(0,1)}\iint_{\R^2} |G(\ve{x};\ve{s})|\,ds_2\,ds_3\Bigr)\cdot
\int_0^1\!\!\sup_{(s_2,s_3)\in\R^2}
\norm{\pt_{\xi_k} \tilde h(\ve{s};\cdot)}{1\,;\Omega}\,ds_1\,,\\
\norm{\pt_{x_k}\bar v(\ve{x};\cdot)}{1\,;\Omega}
&\le \Bigl(\,\sup_{\ve{s}\in\Omega}\norm{G(\ve{s};\cdot)}{1\,;\Omega}.\Bigr)\cdot
\norm{\pt_{x_k}\bar h(\ve{x};\cdot)}{1\,;\Omega}.
\end{align*}
So, in view of \eqref{G_L1}, to prove \eqref{eq:thm:G_eta}, it remains to show that
\[
\int_0^1\!\!\sup_{(x_2,x_3)\in\R^2}\norm{\pt_{\xi_k} \tilde h(\ve{x};\cdot)}{1\,;\Omega}\,dx_1
\leq C\eps^{-1/2},
\qquad
\norm{\pt_{x_k}\bar h(\ve{x};\cdot)}{1\,;\Omega}
\le C\eps^{-1/2}.
\]
These two bounds follow from the definitions \eqref{tilde_bar_h}, \eqref{R_def}
of $\tilde h$ and $\bar h$, which imply that
\begin{align*}
|\pt_{\xi_k} \tilde h(\ve{x};\ve\xi)|
&\le |R\,\partial^2_{x_1 \xi_k}\tilde G|
+C\bigl( |\pt_{x_1}\tilde G|+|\pt_{\xi_k}\tilde G|\bigr)
+|\pt_{\xi_k} \tilde\phi|,\\
|\pt_{x_k}\bar h(\ve{x};\ve\xi)|
&\le |R\,\pt^2_{\xi_1 \xi_k}\bar G|
+C\bigl(|\pt_{\xi_1}\bar G|+ |\pt_{x_k}\bar G|\bigr)
+|\pt_{x_k}\bar\phi|,
\end{align*}
combined with the bounds \eqref{tilde_bar_w} for $\bar\phi$, $\tilde\phi$,
the bounds \eqref{G_tilde_eta}, \eqref{G_tilde_eta_x} for $\tilde G$
and the bounds \eqref{bar_G_xi}, \eqref{bar_G_y} for $\bar G$.
Thus we have shown \eqref{eq:thm:G_eta}.
\smallskip
(ii)
Next we proceed to obtaining
the assertions \eqref{eq:thm:G_xi},
\eqref{eq:thm:G_xixi} and \eqref{eq:thm:G_etaeta}.
We claim that to get these bounds, it suffices to show that
\begin{subequations}\label{desired}
\begin{align}
\mathcal{V}
:=\max_{k=2,\,3}\,\,\sup_{\ve{x}\in\Omega}\norm{\pt^2_{\xi_k}\bar v(\ve{x};\cdot)}{1\,;\Omega}
&\le C(\eps^{-1}+\eps^{-1/2}\, \mathcal{W})
\label{bar_v_eta_eta}\\
\mathcal{W}
:=\sup_{\ve{x}\in\Omega}\norm{\partial_{\xi_1} G(\ve{x};\cdot)}{1\,;\Omega}
&\le C(1+|\ln\eps|+\eps\mathcal{V}),\label{mathcal_G}\\
\sup_{\ve{x}\in\Omega}\|\pt^2_{\xi_1}\bar v(\ve{x};\cdot)\|_{1\,;\Omega}
&\le C\,\eps^{-1}(1+\eps\mathcal{V}).\label{bar_v_xi_xi}
\end{align}
\end{subequations}
Indeed, there is a sufficiently small constant $c_*$ such that
for $\eps\leq c_*$, combining the bounds \eqref{bar_v_eta_eta},\,\eqref{mathcal_G},
one gets $\mathcal{W}\le C(1+|\ln\eps|)$, which is identical with
\eqref{eq:thm:G_xi}.
Then \eqref{bar_v_eta_eta} implies that $\mathcal{V}\le C\eps^{-1}$,
which, combined with \eqref{bar_G_eta_eta_2},
yields \eqref{eq:thm:G_etaeta}.
Finally, $\mathcal{V}\le C\eps^{-1}$ combined with
\eqref{bar_v_xi_xi} and then
\eqref{bar_G_xi_xi_2} yields \eqref{eq:thm:G_xixi}.
In the simpler non-singularly-perturbed case of $\eps>c_*$,
by imitating part (i) of this proof, one obtains $\mathcal{W}\le C_1$,
where $C_1$ depends on $c_*$.
Combining this bound with \eqref{bar_v_eta_eta} and \eqref{bar_v_xi_xi},
we again get \eqref{eq:thm:G_xi}, \eqref{eq:thm:G_xixi} and
\eqref{eq:thm:G_etaeta}.
We shall obtain \eqref{bar_v_eta_eta} in part (iii)
and \eqref{mathcal_G} with \eqref{bar_v_xi_xi} in part (iv) below.
\smallskip
(iii)
To get \eqref{bar_v_eta_eta},
it suffices to set $k=2$ and
consider $\bar V:=\pt^2_{\xi_2}\bar v$ (as $\pt^2_{\xi_3}\bar v$
is estimated similarly).
The problem \eqref{prob_bar_v} for $\bar v$ implies that
\begin{gather}\label{bar_V}
L^*_{\ve\xi}\bar V(\ve{x};\ve\xi)=\bar H(\ve{x};\ve\xi)
\;\;\mbox{for}\;\ve\xi\in\Omega,\quad
\bar V(\ve{x};\ve\xi)=0\;\;\mbox{for}\;\ve\xi\in\partial\Omega.
\end{gather}
The homogeneous boundary conditions $\pt^2_{\xi_2}\bar v\bigr|_{\xi_m=0,1}=0$
in \eqref{bar_V}
for $m=1,\,3$
immediately follow from $\bar v\bigr|_{\xi_m=0,1}=0$.
The homogeneous boundary conditions on the boundary edges
$\xi_2=0,1$ are obtained as follows. As $\bar v\bigr|_{\xi_2=0,1}=0$ so
$\pt_{\xi_1}\bar v\bigr|_{\xi_2=0,1}=\pt^2_{\xi_m}\bar v\bigr|_{\xi_2=0,1}=0$,
where again $m=1,\,3$.
Combining this with $\bar h\bigr|_{\xi_2=0,1}=0$
(for which, in view of Remark~\ref{rem:strip_to_square}, we used \eqref{eq:bar_phi_boundary})
and the differential equation for $\bar v$ at $\xi_2=0,1$, one
finally gets
$\pt^2_{\xi_2}\bar v\bigr|_{\xi_2=0,1}=0$.
For the right-hand side $\bar H$ in \eqref{bar_V}, a calculation shows
that $\bar H=\bar H(\ve{x};\ve\xi)=\pt_{\xi_2}\bar h_1+\bar h_2$
with $\bar h_{1}=\bar h_{1}(\ve{x};\ve\xi)$ and
$\bar h_{2}=\bar h_{2}(\ve{x};\ve\xi)$ defined by
\[
\bar h_1:=\partial_{\xi_k}\bar h
-2\pt_{\xi_k} a(\ve\xi)\cdot\pt_{\xi_1}\bar v,
\qquad
\bar h_2:=\pt^2_{\xi_k} a(\ve\xi)\cdot\pt_{\xi_1}\bar v
-2\pt_{\xi_k} b(\ve\xi)\cdot\pt_{\xi_k}\bar v
-\pt^2_{\xi_k} b(\ve\xi)\cdot\bar v,
\]
with $k=2$.
Here we used
$ \pt^2_{\xi_k}[a\,\pt_{\xi_1}\bar v]
=a\,\pt_{\xi_1}\! \bar V+2\pt_{\xi_k} a \,\pt^2_{\xi_1\xi_k}\bar v
+\pt^2_{\xi_k} a\,\pt_{\xi_1}\bar v
=a\,\pt_{\xi_1}\! \bar V+\pt_{\xi_k}[2\,\pt_{\xi_k} a \,\pt_{\xi_1}\bar v]
-\pt^2_{\xi_k} a\,\pt_{\xi_1}\bar v$
and
$\pt^2_{\xi_k}[b\bar v]=b \bar V+2\,\pt_{\xi_k} b \,\pt_{\xi_k}\bar v+\pt^2_{\xi_k} b\,\bar v$.
(Note that $\bar H$ is understood in the sense of distributions; see Remark \ref{rem_H} below.)
Now, applying the solution representation formula \eqref{eq:sol_adj} to problem \eqref{bar_V},
and then integrating the term with $\bar h_1$ by parts, yields
\[
\bar V(\ve{x};\ve\xi)
=\iiint_\Omega\!\bigl[-\pt_{s_2} G(\ve{s};\ve\xi)\,\bar h_1(\ve{x};\ve{s}
G(\ve{s};\ve\xi)\,\bar h_2(\ve{x};\ve{s})\bigr]\,d\ve{s},
\]
(for the validity of the above integration by parts we again refer
to Remark~\ref{rem_H}).
As \eqref{eq:thm:G_eta} implies $\sup_{\ve{s}\in\Omega}\norm{\pt_{s_2} G(\ve{s};\cdot)}{1\,;\Omega}\le C\eps^{-1/2}$,
while \eqref{G_L1} implies $\sup_{\ve{s}\in\Omega}\|G(\ve{s};\cdot)\|\le C$,
imitating the argument used in part (i) of this proof yields
\[
\norm{\pt^2_{\xi_2}\bar v(\ve{x};\cdot)}{1\,;\Omega}
= \norm{\bar V(\ve{x};\cdot)}{1\,;\Omega}
\leq C\bigl( \eps^{-1/2}\norm{\bar h_1(\ve{x};\cdot)}{1\,;\Omega}
+\norm{\bar h_2(\ve{x};\cdot)}{1\,;\Omega}\bigr).
\]
So to get our assertion \eqref{bar_v_eta_eta}, it remains to show that
\begin{equation}\label{h12_bound}
\norm{\bar h_{1}(\ve{x};\cdot)}{1\,;\Omega}+\norm{\bar h_{2}(\ve{x};\cdot)}{1\,;\Omega}\le C(\eps^{-1/2}+\mathcal{W}).
\end{equation}
To check this latter bound, note that
$|\bar h_1|+|\bar h_2|
\le C(|\pt_{\xi_k}\bar h|+|\pt_{\xi_1}\bar v|+|\pt_{\xi_k}\bar v|+|\bar v|)$
with $k=2$.
Note also that
\[
\norm{\bar v(\ve{x};\cdot)}{1,1\,;\Omega}
\le C(\eps^{-1/2}+\mathcal{W})+\norm{\bar G(\ve{x};\cdot)}{1,1\,;\Omega},
\]
where we employed $\bar v= G-\bar G$ and then the bounds \eqref{G_L1},
\eqref{eq:thm:G_eta} and the definition \eqref{mathcal_G} of $\mathcal{W}$ for $G$.
Combining these two observations with
\[
|\pt_{\xi_k}\bar h(\ve{x};\ve\xi)|
\le |R\,\pt^2_{\xi_1\xi_k} \bar G|
+C\bigl(|\pt_{\xi_1}\bar G|+|\pt_{\xi_k}\bar G|+|\bar G|\bigr)
+|\pt_{\xi_k}\bar \phi|,
\qquad k=2,
\]
(where we used \eqref{bar_h}, \eqref{R_def}), and then with the bounds
\eqref{G_bar_t_1}--\eqref{bar_G_xi_eta} for $\bar G$, and the bound
\eqref{tilde_bar_w} for $\bar\phi$, one gets the required estimate~\eqref{h12_bound}.
Thus \eqref{bar_v_eta_eta} is established.
\smallskip
(iv)
To prove \eqref{mathcal_G} and \eqref{bar_v_xi_xi}, rewrite the problem \eqref{prob_bar_v} as a two-point
boundary-value problem in $\xi_1$, in which $\ve{x}$, $\xi_2$ and $\xi_3$ appear
as parameters, as follows
\begin{gather}\label{eq:ode_barv}
[-\eps\pt^2_{\xi_1}+a(\ve\xi)\pt_{\xi_1}]\,\bar v(\ve{x};\ve\xi)
=\bar{\bar h}(\ve{x};\ve\xi)
\quad\mbox{for}\;\xi_1\in(0,1),
\qquad \bar v(\ve{x};\ve\xi)\bigr|_{\xi_1=0,1}=0,
\end{gather}
where
\begin{gather}\label{bar_bar_h}
\bar{\bar h}(\ve{x};\ve\xi)
:=\bar h(\ve{x};\ve\xi)
+\eps\,\bigl[\pt^2_{\xi_2}\bar v(\ve{x};\ve\xi)
+\pt^2_{\xi_3}\bar v(\ve{x};\ve\xi)\bigr]
-b(\ve\xi)\,\bar v(\ve{x};\ve\xi).
\end{gather}
Consequently, one can represent $\bar v$ via the Green's function
$\Gamma=\Gamma(\xi_1,\xi_2,\xi_3;s)$
of the one-dimensional operator $[-\eps\pt^2_{\xi_1}+a(\ve\xi)\pt_{\xi_1}]$.
Note that $\Gamma$, for any fixed $\xi_2$, $\xi_3$ and $s$, satisfies the equation
$[-\eps\pt^2_{\xi_1}+a(\ve\xi)\pt_{\xi_1}]\Gamma(\ve\xi;s)=\delta(\xi_1-s)$
and the boundary conditions $\Gamma(\ve\xi;s)\bigr|_{\xi_1=0,1}=0$.
Note also that
\begin{gather}\label{G_1d}
\int_{0}^{1}\!\! |\pt_{\xi_1} \Gamma(\ve\xi;s)|\,d\xi_1
\le 2\alpha^{-1}
\end{gather}
\cite[Lemma 2.3]{And_MM02}; see also
\cite[(I.1.18)]{RST08}, \cite[(3.10b) and Section~3.4.1.1]{Linss10}.
The solution representation for $\bar v$ via $\Gamma$ is given by
\[
\bar v(\ve{x};\ve\xi)=\int_0^1\! \Gamma(\ve\xi;s)\, \bar{\bar h}(\ve{x};s,\xi_2,\xi_3)\,ds.
\]
Applying $\pt_{\xi_1}$ to this representation yields
\[
\norm{\pt_{\xi_1} \bar v(\ve{x};\cdot)}{1\,;\Omega}
\le\left(\sup_{(s,\xi_2,\xi_3)\in \Omega}\int_{0}^{1}\!\! |\pt_{\xi_1} \Gamma(\ve\xi;s)|d\xi_1\right)
\cdot\bignorm{\bar{\bar h}(\ve{x};\cdot)}{1\,;\Omega}.
\]
In view of \eqref{G_1d}, we now have
$\norm{\pt_{\xi_1} \bar v}{1\,;\Omega} \le 2 \alpha^{-1} \norm{\bar{\bar h}}{1\,;\Omega}$.
Note that the differential equation \eqref{eq:ode_barv} for $\bar v$
implies that
$\eps\norm{\pt_{\xi_1}^2\bar v}{1;\Omega}
\leq C(\norm{\pt_{\xi_1}\bar v}{1;\Omega}+
\norm{\bar{\bar h}}{1;\Omega})$.
So, furthermore, we get
\[
\| \pt_{\xi_1} \bar v\|_{1\,;\Omega}
+\eps\| \pt^2_{\xi_1} \bar v\|_{1\,;\Omega} \le C \|\bar{\bar h}\|_{1\,;\Omega}.
\]
As $G=\bar v+\bar G$ and we have the bound \eqref{bar_G_xi} for $\pt_{\xi_1} \bar G$,
to obtain the desired bounds \eqref{mathcal_G} and \eqref{bar_v_xi_xi}, it remains to show that
$\norm{\bar{\bar h}(\ve{x};\cdot)}{1\,;\Omega}\le C+\eps\mathcal{V}$.
Furthermore, the definitions \eqref{bar_bar_h} of $\bar{\bar h}$
and \eqref{bar_v_eta_eta} of $\mathcal{V}$,
imply that it now suffices to prove the two estimates
\begin{gather}\label{two_bounds}
\norm{\bar v(\ve{x};\cdot)}{1\,;\Omega}\le C,
\qquad
\norm{{\bar h}(\ve{x};\cdot)}{1\,;\Omega}\le C.
\end{gather}
The first of them follows from $\bar v=G-\bar G$
combined with \eqref{G_L1} and \eqref{G_bar_t_1}.
The second is obtained from the definition \eqref{bar_h} of ${\bar h}$
using \eqref{bar_G_y} for $\norm{R\pt_{\xi_1}\bar G}{1\,;\Omega}$,
\eqref{G_bar_t_1} for $\norm{\bar G}{1\,;\Omega}$
and \eqref{tilde_bar_w} for $\norm{\bar \phi}{1\,;\Omega}$.
This completes the proof of \eqref{mathcal_G} and \eqref{bar_v_xi_xi}, and thus of
\eqref{eq:thm:G_xi}, \eqref{eq:thm:G_xixi} and \eqref{eq:thm:G_etaeta}.
\smallskip
(v)
We now focus on the remaining
assertion \eqref{eq:thm:G_grad},
again rewrite the problem \eqref{prob_bar_v} as
\[
[-\eps\laplace_{\ve\xi}+1]\,\bar v(\ve{x};\ve\xi)
={\bar h}_0(\ve{x};\ve\xi)\quad\mbox{for}\;\ve\xi\in\Omega,
\qquad \bar v(\ve{x};\ve\xi)\bigr|_{\pt\Omega}=0,
\]
where
\begin{gather}\label{bar_bar_h_new}
{\bar h}_0(\ve{x};\ve\xi)
:= \bar h(\ve{x};\ve\xi)
-a(\ve\xi)\,\pt_{\xi_1} \bar v(\ve{x};\ve\xi)
+[1-b(\ve\xi)]\,\bar v(\ve{x};\ve\xi).
\end{gather}
We shall represent $\bar v$ via the Green's function $\Psi$
of the two-dimensional self-adjoint operator
$[-\eps\laplace_{\ve\xi}+1]$.
Note that $\Psi=\Psi(\ve{s};\ve\xi)$, for any fixed $\ve{s}$, satisfies the equation
$[-\eps\laplace_{\ve\xi}+1]\Psi(\ve{s};\ve\xi)=\delta(\ve\xi-\ve{s})$,
and also the boundary conditions $\Psi(\ve{s};\ve\xi)\bigr|_{\ve\xi\in\pt\Omega}=0$.
Furthermore, for any ball $B(\ve{x}';\rho)$ of radius $\rho$ centred at any
$\ve{x}'$, we cite the estimate \cite[(3.5b)]{CK09}
\begin{gather}\label{G_2d}
|\Psi(\ve{s};\cdot)|_{1,1\,;B(\ve{x}';\rho)\cap\Omega} \le C\eps^{-1}\rho.
\end{gather}
The solution representation for $\bar v$ via $\Psi$ is given by
\[
\bar v(\ve{x};\ve\xi)
=\iiint_\Omega \Psi(\ve{s};\ve\xi)\,{\bar h}_0(\ve{x};\ve{s})\,d\ve{s}.
\]
Applying $\pt_{\xi_m}$, $m=1,\,2,\,3$ to this representation yields
\begin{gather}\label{ber_v_new}
|\bar v(\ve{x};\cdot)|_{1,1\,;B(\ve{x}';\rho)\cap\Omega}
\le \Bigl(\sup_{\ve{s}\in\Omega}|\Psi(\ve{s};\cdot)|_{1,1\,;B(\ve{x}';\rho)\cap\Omega}\Bigr)\cdot
\norm{{\bar h}_0(\ve{x};\cdot)}{1\,;\Omega}.
\end{gather}
To estimate $\norm{{\bar h}_0}{1\,;\Omega}$, recall that
it was shown in part (iv) of this proof that
$\norm{\pt_{\xi_1} \bar v}{1\,;\Omega} \le 2 \alpha^{-1} \norm{\bar{\bar h}}{1\,;\Omega}$
and
$\norm{\bar{\bar h}(\ve{x};\cdot)}{1\,;\Omega}\le C+\eps\mathcal{V}$,
and in part (ii) that $\mathcal{V}\le C\eps^{-1}$.
Consequently $\norm{\pt_{\xi_1} \bar v}{1\,;\Omega} \le C$.
Combining this with \eqref{bar_bar_h_new} and \eqref{two_bounds}
yields $\norm{{\bar h}_0}{1\,;\Omega}\le C$.
In view of \eqref{ber_v_new} and \eqref{G_2d},
we now get $|\bar v|_{1,1\,;B(\ve{x}';\rho)\cap\Omega}\le C\eps^{-1}\rho$,
which, combined with \eqref{bar_G_1_1_B},
immediately gives the final desired bound \eqref{eq:thm:G_grad}.
\end{proof}
\begin{rem}\label{rem_H}
Note that the term $\pt^2_{\xi_k}\bar h$ in $\bar H$, where $k=2,\,3$,
has such a singularity at $\ve\xi=\ve{x}$
that it is not absolutely integrable in $\Omega$. So $\bar H$ and
the differential equation in \eqref{bar_V} are understood in the sense of distributions
\cite[Chapters 1,\,3]{Griffel}. In particular $\pt^2_{\xi_k}\bar h$ is a generalised
$\xi_k$-derivative of the regular function $\pt_{\xi_k}\bar h$.
\end{rem}
\bibliographystyle{plain}
|
\section{Introduction}
F-theory provides a promising starting point for connecting string theory
with particle phenomenology. In particular, Grand Unified Theories (GUTs)
can be elegantly realized in this setup \cite{BHVI, BHVII, DWI, DWII} (see
\cite{Heckman:2010bq, Weigand:2010wm} for reviews). Such scenarios feature higher
dimensional GUT theories where the GUT group is localized on a seven-brane,
and matter is trapped at the intersection of the GUT brane with additional
flavor seven-branes. Interactions are localized at common intersections of the
matter fields. The F-theory approach to particle physics is promising because
while it is flexible enough to accommodate the basic features of realistic
particle phenomenology, it is also rigid enough to impose non-trivial
constraints on potential scenarios of physics beyond the Standard Model (SM).
One example of a such a constraint is that generating
a large top quark mass requires the existence of
exceptional (e.g. E-type) symmetry enhancement points in the geometry
\cite{BHVI, DWI}. Another example is that in geometrically minimal
realizations of flavor hierarchies, both the CKM matrix and the mass matrices
are in relatively close accord with observation \cite{HVCKM, HVCP,
FGUTSNC, BHSV}. Certain supersymmetry (SUSY) breaking scenarios such as a ``PQ
deformation'' of minimal gauge mediation can also be accommodated
\cite{HVGMSB} in which the messenger fields localize on curves of the
geometry. Combining this with other geometric conditions
imposes constraints missing from purely low energy considerations; for example, in
many models the representation content of the messengers is fixed by the
available ways to unfold a singularity \cite{EPOINT}. This is
quite economical and constitutes a self-contained package. Nevertheless, given
the existence of an entire landscape of possible closed string vacua, it is
important to investigate possible well-motivated extensions of this framework.
A robust feature of this setup is that the possible ways to extend such GUT
models by additional seven-brane intersections are rather limited. Aside from
seven-branes, the main ingredients generically present in compactifications of
F-theory are probe D3-branes. Such D3-branes fill 3+1 noncompact spacetime
dimensions and sit at points of the internal geometry. Their presence in a globally complete
compactification is often required in order to cancel tadpoles. Moreover, they
are also locally attracted to the E-type Yukawa points of the visible sector
\cite{Funparticles}. It is therefore natural to ask what type of extra sectors
are realized on such probe theories. At a generic internal point of the
geometry, this extra sector is not terribly interesting. For example, a single
D3-brane realizes $U(1)$ $\mathcal{N}=4$ Super Yang-Mills theory. In this
theory, the vevs of the three chiral fields control the local position of the
D3-brane in the three complex-dimensional internal geometry.
However, there is a distinguished point in F-theory phenomenology where
intersecting seven-branes realize an $E_{6,7,8}$ exceptional symmetry. In geometrically
minimal constructions, a single point of $E_8$ is responsible for generating all
of the relevant visible sector Yukawa couplings \cite{BHSV, EPOINT}. Since the D3-brane is attracted to such Yukawa
points (see \cite{Martucci, FGUTSNC, Funparticles}), it is natural to
ask what happens if D3-branes reside at (or very near) such a point. It was
with this motivation in mind that these theories were proposed as possible
quasi-hidden sectors for F-theory GUTs in \cite{Funparticles} which were
further studied in \cite{FCFT, D3gen}.
The dominant couplings between the probe and Standard Model (SM) degrees of freedom
have been studied in \cite{Funparticles, TBRANES}. Some of the states of the
probe are charged under $SU(5)_{GUT}$ and therefore communicate via gauge
interactions with the visible sector. Another source of probe/MSSM couplings is
via F-terms. These F-terms primarily couple the probe to the Higgs fields and
third generation of matter fields. This is because in the flavor hierarchy
scenario of \cite{HVCKM, BHSV, EPOINT}, these are the modes with maximum
overlap with the Yukawa point, which is also the location of the D3-brane.
In the limit where one neglects the effects of the Standard Model fields, it
was found in \cite{FCFT} that such probe theories realize a new class of
strongly interacting $\mathcal{N} = 1$ superconformal field theories (SCFTs). These
theories are defined by starting from the $\mathcal{N} = 2$
Minahan-Nemeschansky theories with $E_8$ flavor symmetry \cite{MNI, MNII},
and adding a set of field-dependent mass deformations in the $SU(5)_{\bot}$
factor of $E_{8} \supset SU(5)_{GUT} \times SU(5)_{\bot}$.
The choice of mass deformations is dictated by the local configuration of
intersecting seven-branes. Remarkably, the configurations which lead to
new $\mathcal{N} = 1$ fixed points are precisely those which are of most
relevance for phenomenology.
With an eye towards potential model building applications, in this paper we
take the next step in this analysis by coupling the probe to a dynamical
visible sector. In other words, at the GUT scale $M_{GUT}$, we consider
performing a further deformation of the $\mathcal{N} = 1$ theories studied in
\cite{FCFT}. This deformation induces a flow to
another class of interacting CFTs. Much as in \cite{FCFT}, we can compute
various details of the system such as the scaling dimension of chiral
operators, as well as the running of the visible sector gauge couplings. Although the combined
probe/MSSM system realizes an interacting conformal fixed point, there is a
well-defined sense in which the mixing between the two sectors remains small.
Indeed, we find that the scaling dimensions of the SM chiral fields remains
rather close to their free field values, with anomalous dimensions typically in
the range of $0.0 - 0.1$.
Since the CFT living on the D3-brane includes degrees of freedom charged under
$SU(5)_{GUT}$, these modes must develop a mass to have evaded detection thus
far. This places a lower bound on the CFT breaking scale of order
$M_{\cancel{CFT}} \gtrsim$ TeV. It is natural to ask how to break conformal
invariance. A simple possibility is to have the D3-brane slightly displaced
from the Yukawa point, since the displacement automatically introduces a scale and gives mass to these
$SU(5)_{GUT}$ charged states of the probe sector.
Treating the SM gauge group as a weakly gauged flavor symmetry of the conformal
fixed point, we calculate the contribution of the probe to the running of the SM gauge couplings.
Introducing a characteristic scale $M_{\cancel{CFT}}$ associated with conformal symmetry
breaking, we can view this computation as a threshold correction to the
running of the MSSM gauge couplings which enters at energies $M_{\cancel{CFT}}
<E < M_{GUT}$. We find that although the states of the probe sector fill out
full GUT multiplets, the additional charged degrees of freedom of the CFT
\textit{improve} precision unification of the MSSM (which for typical superpartner masses
otherwise contains an order $\sim 4\%$ mismatch at two loop level) due to their $SU(5)_{GUT}$
breaking couplings to the Higgs sector (see also \cite{Donkin:2010ta}). See
figure \ref{unifyplot} for a depiction of these effects.
\begin{figure}
[ptb]
\begin{center}
\includegraphics[
height=4.6198in,
width=6.1765in
]%
{UnifyPlot.eps}%
\caption{Depiction (not drawn to scale) of threshold corrections to gauge
coupling unification from the D3-brane probe sector (black solid curves). This
is to be compared with two-loop MSSM\ and other contributions with similar
effects (red dashed lines). The GUT\ scale is defined as the scale at which
$\alpha_{1}$ and $\alpha_{2}$ unify. The D3-brane threshold correction enters
at the scale of conformal symmetry breaking, which is the characteristic mass
scale for states of the probe charged under $SU(5)_{GUT}$. Here, we have
indicated that the size of the relative shift in couplings $\delta
\equiv(\alpha_{3}^{-1}-\alpha_{GUT}^{-1})/\alpha_{GUT}^{-1}$ is positive for
two-loop effects and negative for the D3-brane threshold. Balancing these
effects improves precision unification.}%
\label{unifyplot}%
\end{center}
\end{figure}
It is also natural to ask what effect the D3-brane has on the visible sector at energies
below $M_{\cancel{CFT}}$. At this point there are two possibilities, depending on whether or not
SUSY breaking takes place on the D3-brane hidden sector. If it
does not, the story is rather simple, and the only remaining effects may
include some light remnants which interact via kinetic mixing with
the visible sector \cite{Funparticles, D3gen}.
If, however, SUSY breaking takes place on the D3-brane, the pattern of soft supersymmetry breaking terms in the visible sector may be strongly affected.
Along these lines, the most natural thing to assume is that the D3-brane position mode
which breaks conformal invariance also develops a SUSY breaking vev. Assuming this, the $SU(5)_{GUT}$ charged mediator states from
the probe to the visible sector play the role of messenger fields in a minimal
gauge mediation scenario, leading to a very predictive superpartner mass
spectrum. The dominant coupling to the third generation and Higgs fields
induces further corrections to the gauge mediated spectrum. These additional
contributions can induce additional third generation/Higgs sector
soft masses and large $A$-terms, as well as $\mu$ and $B \mu$-terms.
The organization of this paper is as follows. In section \ref{sec:QUASI} we
review the basic features of the visible sector in F-theory GUTs, and the
precise definition of the probe sectors we study in the remainder of this
paper. In section \ref{sec:FIXED} we study the infrared fixed points
associated with coupling the MSSM to the probe. Using these results, in
section \ref{sec:UNIFY} we demonstrate that for appropriate threshold scales,
the presence of the probe can actually improve precision unification. We next
turn in section \ref{sec:SUSY} to the potential role of the probe as a sector
for supersymmetry breaking and transmission to the visible sector. Section
\ref{sec:CONC} contains our conclusions. A brief review of probes of T-Branes
is given in Appendix A, and a collection of additional monodromy scenarios is collected
in Appendix B.
\section{Quasi-Hidden Sectors From F-theory}
\label{sec:QUASI}
In this section we review the basic setup we shall be considering. The main
idea is that the Standard Model is realized via a configuration of
intersecting seven-branes in F-theory, and that D3-branes constitute an
additional sector which can interact non-trivially with this system.
Throughout, we shall work in terms of a limit where four-dimensional \ gravity
has been decoupled, so that $M_{pl}^{(4d)}\rightarrow\infty$. This
approximation is especially well-justified in systems where we focus on just
the local intersections of non-compact seven-branes probed by a D3-brane which
is pointlike in the internal directions. We now turn to a more precise
characterization of the system we shall study.
\subsection{Intersecting Seven-Branes and the Visible Sector}
F-theory is a non-perturbative formulation of IIB\ string theory in which the
axio-dilaton $\tau=C_{0}+i/g_{s}$ can attain order one values, which moreover
can have non-trivial position dependence on the internal directions. Going to
strong coupling provides a way to realize intersecting brane configurations
with E-type symmetries. E-type symmetries are quite important in string-based
GUT\ models, thus suggesting F-theory GUTs as a natural framework for string
model building \cite{BHVI,BHVII,DWI,DWII}. The profile of the
dilaton is, in minimal Weierstrass form, dictated by the equation:
\begin{equation}
y^{2}=x^{3}+f(z,z_{1},z_{2})x+g(z,z_{1},z_{2}).
\end{equation}
Here, $z$, $z_1$ and $z_2$ define three local complex coordinates for the threefold base of
any F-theory compactification. The modular parameter of this elliptic curve is the axio-dilaton $\tau_{IIB}$.
The location of the intersecting seven-branes is then given by the
discriminant
\begin{equation}
\Delta\equiv4f^{3}+27g^{2}=0.
\end{equation}
In such models, the visible sector is realized on a configuration of
interecting seven-branes. The local intersections of such seven-branes can be
modelled in terms of eight-dimensional Super Yang-Mills theory with gauge
group $G$. We denote by $z_{1}$ and $z_{2}$ two local complex coordinates
parameterizing the internal worldvolume of the parent theory. The breaking
pattern of $G$ then dictates the locations of localized matter, as well as the
interactions between this matter \cite{KatzVafa, BHVI, DWIII, TBRANES}. For the purposes of
GUT\ model building, it is most convenient to work in terms of $E_{8}$ gauge
theory. The particular choice of a breaking pattern in a local patch of the
geometry is dictated by an adjoint-valued $(2,0)$ form which we denote by
$\Phi$. This $\Phi$, along with the internal gauge fields of the seven-branes,
satisfy a coupled system of F- and D-term equations. For example, to
realize an $SU(5)$ F-theory GUT, one considers the breaking pattern:%
\begin{equation}
E_{8}\supset SU(5)_{GUT}\times SU(5)_{\bot}%
\end{equation}
where $\Phi$ takes values in the adjoint of $SU(5)_{\bot}$. This
tilting of the seven-brane configuration occurs at an energy scale
which we shall refer to as $M_{\ast}$. The
compactification of the seven-brane is assumed to occur at the GUT\ scale
$M_{GUT}<M_{\ast}$. Numerically, we have $M_{\ast}\sim10^{17}$ GeV and
$M_{GUT}\sim2\times10^{16}$ GeV \cite{BHVII}.
The particular choice of $\Phi$ dictates the phenomenology of the visible
sector. A very convenient way to analyze supersymmetric configurations of
intersecting seven-branes is in terms of \textquotedblleft holomorphic
gauge.\textquotedblright\, This choice amounts to analyzing just the F-term equations
of motion for the system, and effectively complexifying the gauge group to
$G_{%
\mathbb{C}
}$. In an appropriate gauge, one can then locally present $\Phi(z_{1},z_{2})$
as a purely holomorphic expression in the local coordinates $z_{1}$, $z_{2}$
of the parent theory worldvolume coordinates. In realistic F-theory GUT
configurations, the actual seven-brane configuration corresponds to a
\textquotedblleft T-brane\textquotedblright\ \cite{TBRANES}. Locally, this configuration is specified
by the condition that at the origin, $\Phi(0,0)$ is
nilpotent but non-zero. In holomorphic gauge, $\Phi(0,0)$ can then be decomposed
into a direct sum of nilpotent Jordan blocks.
\subsection{Review of Probe D3-Branes}
So far, our discussion has focussed on the properties of the visible sector,
e.g., the seven-branes of the system. An additional well-motivated sector to
consider is one with D3-brane probes of such intersecting brane configurations. Away
from the configuration of seven-brane intersections, the probe D3-brane is
given by $\mathcal{N}=4$ Super Yang-Mills theory with $U(1)$ gauge group. The
worldvolume gauge coupling is the IIB axio-dilaton $\tau_{IIB}=\tau_{D3}$. As
the D3-brane moves closer to the configuration of intersecting seven-branes,
additional modes will become light. This can lead to highly non-trivial
interacting superconformal field theories.
\begin{figure}
[ptb]
\begin{center}
\includegraphics[
height=4.6224in,
width=4.9917in
]%
{DefFlow.eps}%
\caption{Depiction of the various deformations of the D3-brane probe theory.
In a limit where gravity is decoupled, in the UV, the D3-brane probes a
parallel stack of seven-branes. At a scale $M_{\ast}$, the stack of
seven-branes is tilted, inducing a flow to an interacting fixed point, denoted
by T-Brane$\oplus D3$. At a scale $M_{GUT}$, a further deformation is added,
corresponding to coupling to dynamical Standard Model fields. This leads to a
new fixed point, denoted by $SM\oplus$T-Brane$\oplus D3$ }%
\label{defflow}
\end{center}
\end{figure}
It is helpful to organize our discussion according to the energy scales of the
probe system. See figure \ref{defflow} for a depiction of the various energy scales
and corresponding deformations of the probe/MSSM system.
At high energies, when the seven-brane configuration is
parallel, the probe realizes an $\mathcal{N}=2$ theory. The effects of
seven-brane tilting are then reflected in the probe theory as a combination of
F- and D-term $\mathcal{N}=1$ deformations, which are added at the scale
$M_{\ast}$. A further deformation appears at the GUT scale, because this is
the scale at which the fields of the Standard Model become dynamical modes of
the four-dimensional theory. One of the aims of this paper is to show that in
the absence of other effects, these deformations lead to interacting fixed
points for the combined probe/MSSM system.
Throughout this work, we will have occasion to refer to the various string
states connecting the seven-branes and three-branes. We label the visible sector stack
as a $7_{SM}$-brane, and the ``flavor brane'' as a $7_{flav}$-brane. The states in the worldvolume
theory of the three-brane probe will come from the various types of $3-3$ or
$3-7$ strings. We refer to states of the probe sector which are also charged
under the SM gauge group as $3-7_{SM}$ strings, and those $3-7$ strings which
are neutral under $SU(5)_{GUT}$ as $3-7_{flav}$ strings. There are also various types
of $7-7$ strings as well, which are important for realizing the visible sector.
Let us note that this is a slight abuse of terminology, because the actual modes will involve both
weakly coupled strings, as well as $(p,q)$ strings and their junctions
\cite{Gaberdiel:1997ud, Gaberdiel:1998mv, DeWolfe:1998zf, DeWolfe:1998bi}.
The examples which have been most extensively studied in the literature
correspond to worldvolume theories which retain $\mathcal{N}=2$ supersymmetry.
In this case, there is a single stack of parallel seven-branes with gauge
group $G$. This becomes a flavor symmetry of the probe theory. A remarkable
feature of these probe theories is that the associated Seiberg-Witten curve is
given by the same F-theory geometry \cite{Sen:1996vd, Banks:1996nj}. Examples of such $\mathcal{N}=2$ probe
theories include $\mathcal{N}=2$ $SU(2)$ SYM with four flavors \cite{Banks:1996nj}, Argyres-Douglas
fixed points \cite{Argyres:1995jj}, and the Minahan-Nemeschansky theories with exceptional flavor
symmetry \cite{MNI, MNII}. Although the first example is simply a D3-brane probing a $D_4$ seven-brane, the latter two
possibilities are realized by D3-branes probing F-theory singularities of type
$H_{i}$ and $E_{n}$, and these involve intrinsically non-perturbative ingredients.
This means that such $\mathcal{N}=2$ theories are strongly coupled. The moduli
space of the $\mathcal{N}=2$ single D3-brane probe theory is given by a Higgs
branch and a Coulomb branch. These are, respectively, parameterized by the vev of a
dimension two operator $\mathcal{O}$ in the adjoint of the flavor group $G$,
and the vev of a field $Z$ which parameterizes motion normal to the seven-brane. In
addition, there is a decoupled hypermultiplet $Z_{1}\oplus Z_{2}$
parameterizing motion parallel to the seven-brane.
In the weakly coupled example given by a D3-brane probing a $D_{4}$ singularity, we
have $\mathcal{N} = 2$ super Yang-Mills theory with gauge group $SU(2)$ and
hypermultiplets $Q\oplus\tilde{Q}$. In this case, the operator
$\mathcal{O}$ is given by $\mathcal{O}\sim Q\tilde{Q}$ and has
scaling dimension two. In a theory with exceptional flavor
symmetries we lose any description of this operator in terms of elementary
fields, although one can show there are still dimension two operators
parameterizing the Higgs branch.
Tilting the seven-branes at a scale $M_{\ast}$ corresponds in the D3-brane theory
to coupling the position modes $Z_1$ and $Z_2$ to the $\mathcal{N} = 2$ theory.
This coupling breaks breaks part of the $E_8$ flavor symmetry.
Although we find the breaking $E_{8}\rightarrow SU(5)_{GUT} \times SU(5)_{\bot}$ to
be most useful for model building purposes, there is no \textit{a priori}
reason that we could not choose a different breaking pattern. This tilting
gives a mass to some of the $3-7_{flav}$ strings, and is reflected in the
probe theory as a superpotential deformation:
\begin{equation}
\delta W_{tilt}=\,\mathrm{Tr}_{G}\left( \Phi(Z_{1},Z_{2})\cdot\mathcal{O}%
\right) , \label{trphio}%
\end{equation}
where $\Phi(Z_{1},Z_{2})$ is valued in the adjoint of the global symmetry of
the probe theory. Choices of $\Phi$ for which $[\Phi,\Phi^{\dagger}]\neq0$
will break half of the supersymmetries, leaving $\mathcal{N}=1$ in the probe
theory. Let us note that in most realistic T-brane configurations, $[\Phi
,\Phi^{\dagger}]\neq0$ since the constant part of $\Phi$ is a non-zero
nilpotent matrix.
As found in \cite{FCFT}, superpotentials of the form (\ref{trphio}) often lead
to new strongly coupled $\mathcal{N}=1$ SCFTs in the IR. If these SCFTs
descend from $\mathcal{N}=2$ theories with exceptional global symmetries, they
will not have a known UV Lagrangian description. As emphasized in \cite{FCFT},
this does not mean that such theories are totally inaccessible to study. Since
we know the global symmetries and charges of various operators, we can still
use $a$-maximization \cite{Intriligator:2003jj} to compute the scaling dimensions of chiral primary
operators. The resulting data provides a non-trivial consistency check that such
$\mathcal{N}=1$ SCFTs exist in the IR.
Compactifying the seven-brane theory leads to an additional deformation of the
probe theory. As a first approximation, we work in the limit where the gauge
theory of the Standard Model remains as a flavor symmetry of the D3-brane
sector, but where the Standard Model fields localized on curves are dynamical.
Let us note that this is a justified approximation in F-theory GUTs, because
the gauge fields propagate over a four-dimensional worldvolume, while the
matter fields can localize on compact two-dimensional subspaces.
At the level of F-terms, the effects of compactifying the matter curves of the
system means that an additional deformation is added at the GUT scale:%
\begin{equation}
\delta W_{SM\oplus D3}=\Psi_{R}^{(SM)}\mathcal{O}_{\overline{R}}+W_{MSSM}
\label{SMdef}%
\end{equation}
for a Standard Model field $\Psi_{R}^{(SM)}$ in a representation $R$ of the
Standard Model gauge group, and operators $\mathcal{O}_{\overline{R}}$ in the
dual representation. Additionally, $W_{MSSM}$ consists of the leading order
cubic terms between the zero modes:%
\begin{equation}
W_{MSSM}=\lambda_{ij}^{(l)}H_{d}L^{i}E^{j}+\lambda_{ij}^{(d)}H_{d}Q^{i}%
D^{j}+\lambda_{ij}^{(u)}H_{u}Q^{i}U^{j}.
\end{equation}
Here, $i,j=1,2,3$ are indices running over the generations of Standard Model
fields. We neglect the contribution from the $\mu$-term and other effects
which are associated with supersymmetry breaking, as these are expected to be
induced at far lower energy scales.
The precise form of the couplings $\Psi_{R}^{(SM)}\mathcal{O}_{\overline{R}}$
depends on the details of the local profile of the matter fields, which is in
turn determined by the choice of tilting parameter $\Phi$. We review the form of these
couplings in Appendix A. See \cite{TBRANES} for further discussion.
One qualitative feature of these couplings is that in models where all three
generations localize on a single matter curve, the dominant coupling to the
probe sector will be via those modes which have maximal overlap with the
Yukawa point. Using the model of flavor physics developed in \cite{HVCKM,
BHSV, HVCP, FGUTSNC}, this maximal overlap means that the third generation (e.g. the heaviest
modes) of the Standard Model will be the ones which dominantly couple to the probe sector.
Since the Higgs up and Higgs down also naturally localize on distinct matter
curves, they will also have order one profiles at the Yukawa
point. Thus, the primary couplings are to the third generation, and the Higgs
fields. There will be some coupling to the first and second generation matter
fields, though these couplings are expected to be suppressed by powers of
$\alpha_{GUT}$. In this paper we neglect such couplings, since they are a
small correction to the basic features considered here. In this case, the
superpotential deformation can be written as:%
\begin{equation}
\delta W_{SM\oplus D3}=H_{u}\mathcal{O}_{H_{u}}+H_{d}\mathcal{O}_{H_{d}%
}+\overline{5}_{M}\mathcal{O}_{\overline{5}_{M}}+10_{M}\mathcal{O}_{10_{M}%
}+W_{MSSM}%
\end{equation}
in the obvious notation.
Gauge field interactions constitute another source of interaction between the
probe and the Standard Model. For example, since the probe contains states
charged under the Standard Model gauge group, these modes will couple to the
corresponding gauge fields, affecting the running of couplings. Additionally,
kinetic mixing between the probe and Standard Model is generically expected
\cite{Funparticles}. All of these couplings lead to a very rich quasi-hidden
sector. One of our aims in this paper will be to quantify some details of the mixed
probe/MSSM system.
\section{Fixed Points of the Coupled Probe/MSSM System\label{sec:FIXED}}
In the previous section we reviewed the basic setup for the system described
by a strongly coupled D3-brane probing the Standard Model. In this section we
consider a particular idealization where we do not include effects from breaking
either the conformal symmetry or supersymmetry. Further, we work in the limit where the
SM gauge group can be treated as a flavor symmetry so that it is
only weakly gauged. However, we do allow the Standard Model fields lcoalized on
curves to be dynamical. Geometrically, this corresponds to a limit where some
of the matter curves are compact, while the full GUT\ seven-brane is non-compact.
Under these circumstances, we provide evidence that the coupled probe/MSSM
systems will flow to non-trivial $\mathcal{N}=1$ superconformal
fixed points. The precise fixed point in question is controlled by the choice
of UV deformation, which we summarize as:%
\begin{equation}
\delta W_{UV}=\delta W_{tilt}+\delta W_{SM\oplus D3}\text{.}%
\end{equation}
This depends on the choice of $\Phi$, as well as the zero mode content on the
matter curves. Perhaps surprisingly, we find that not only does the
coupled probe/MSSM system admit such fixed points, but that the operators of
the Standard Model develop only \textit{small} anomalous dimensions.
Due to the small shift in the scaling dimensions of Standard Model fields, there is a well-defined sense
in which the Standard Model degrees of freedom retain their identity. This is
important, because it means that once we include the effects of CFT breaking,
the identity of the Standard Model states remains roughly the same as in the
UV description in terms of intersecting seven-branes. The low amount of mixing is
further corroborated by the computable effect of the probe on the running of the SM gauge couplings.
Indeed, we find only a mild (but irrational) shift to the usual one-loop beta functions.
In much of this paper, we shall focus on the effects of a single probe D3-brane, though we shall
also consider the case of two D3-branes as well. Indeed, adding a large number of D3-branes
induces a much larger threshold correction to the running of the SM gauge couplings. This in
turn can induce a Landau pole before the GUT scale. In the context of F-theory GUTs which
admit a decoupling limit \cite{BHVII, HVGMSB, HeckVerlinde}, there is actually a more stringent requirement
that the zero mode content must preserve asymptotic freedom of $SU(5)_{GUT}$.
This limits the amount of additional matter which can be added below the GUT scale.
To get a rough sense of the amount of allowed extra matter at low energies,
we can consider the beta function associated with three generations of
chiral matter, one vector-like pair of $5_{H} \oplus \overline{5}_{H}$, and $\delta b_{SU(5)}$
additional GUT multiplets in the $5 \oplus \overline{5}$. The resulting beta function
is, in our sign conventions:
\begin{equation}
b_{SU(5)} = - 15 + 7 + \delta b_{SU(5)}.
\end{equation}
In other words, if we demand $b_{SU(5)} < 0$, we obtain the condition
$\delta b_{SU(5)} < 8$. In practice, we find that in most probe scenarios, this bound
limits us to one or two probe D3-branes.
The rest of this section is organized as follows. By appealing to the symmetries of our
non-Lagrangian probe theory we determine up to one unfixed parameter the
general form of the infrared R-symmetry. Much as in \cite{FCFT}, this parameter
is then fixed by $a$-maximization \cite{Intriligator:2003jj}. Determining the infrared R-symmetry
allows us to compute the scaling dimensions of chiral primary
operators such as the MSSM chiral superfields. We also show that
in the limit where the Standard Model gauge group is treated as a weakly gauged flavor symmetry,
we can also extract the change in the beta functions from the probe theory. After this, we present in
detail a particular example of a probe theory. This is followed by a summary of various probe scenarios involving
one, as well as two probe D3-branes.
\subsection{Infrared R-Symmetry}
\bigskip In general, knowing the global symmetries of an SCFT is enough to
figure out the R-charges (or equivalently the scaling dimensions) of all
chiral primary operators. The procedure for doing this was first described in
\cite{Intriligator:2003jj}, in which it was discovered that the superconformal
R-symmetry is the one that locally maximizes
\begin{equation}
a_{trial}=\frac{3}{32}\left( 3\,\mathrm{Tr}R_{trial}^{3}-\,\mathrm{Tr}%
R_{trial}\right) , \label{atrial}%
\end{equation}
where $R_{trial}=R_{0}+\sum_{I}s_{I}F_{I}$ is a general linear combination of
an R-symmetry $R_{0}$ and all anomaly-free global symmetries $F_{I}$. Finding
the unique local maximum of (\ref{atrial}) then fixes all the coefficients
$s_{I}$.
Assuming the absence of accidental IR\ symmetries, we know the set of global
symmetries of the SCFTs we study in this work. These are given by the
symmetries of the original $\mathcal{N}=2$ probe theory, and additional
symmetries which act on the Standard Model. The various charges of the probe
sector and Standard Model fields, in the case where we begin with an $E_8$ $\mathcal{N}=2$ theory, are summarized below%
\begin{equation}
\label{tabtab}%
\begin{tabular}
[c]{|c|c|c|c|c|c|}\hline
UV symmetries & $\mathcal{O}_{s}$ & $Z$ & $Z_{1}$ & $Z_{2}$ & $\Psi_{SM}%
$\\\hline
$R_{UV}$ & $\frac{4}{3}$ & $4$ & $\frac{2}{3}$ & $\frac{2}{3}$ & $\frac{2}{3}%
$\\\hline
$J_{\mathcal{N}=2}$ & $-2$ & $12$ & $-1$ & $-1$ & $0$\\\hline
$U(1)_{1}$ & $0$ & $0$ & $+1$ & $0$ & $0$\\\hline
$U(1)_{2}$ & $0$ & $0$ & $0$ & $+1$ & $0$\\\hline
$U(1)_{\Psi}$ & $0$ & $0$ & $0$ & $0$ & $+1$\\\hline
$T_{3}$ & $s$ & $0$ & $0$ & $0$ & $s_{\Psi}$\\\hline
\end{tabular}
\ .
\end{equation}
Here, $R_{UV}$ denotes the UV R-symmetry of the theory when treated in terms
of a decoupled $\mathcal{N}=2$ system and a weakly coupled set of free fields
for the Standard Model degrees of freedom. In the context of an $\mathcal{N}%
=2$ superconformal field theory, $R_{UV}$ and $J_{\mathcal{N}=2}$
combine to form an abelian subalgebra of the $SU(2)\times U(1)$ R-symmetry. The unusual $J_{\mathcal{N}=2}$ charge of $Z$ is a result of using the $E_8$ theory as our starting point. We
have also indicated the rephasing symmetries for the decoupled hypermultiplet
$Z_{1}\oplus Z_{2}$. On similar grounds, we have also included a set of
rephasing symmetries for the various Standard Model fields. For each Standard
Model field, there is a corresponding $U(1)_{\Psi}$. We note that although
such symmetries will be broken by the presence of cubic interaction terms in
$W_{MSSM}$, such interaction terms do not affect the IR fixed point since they
are irrelevant in the IR.\footnote{In more
precise terms, we note that the MSSM\ superpotential consists of terms cubic
in Standard\ Model operators, since the $\mu$-term is assumed to be generated
at scales below the CFT\ breaking scale. Given a cubic coupling, the only way
that this term can be maintained as a marginal coupling in the IR is if
all Standard Model fields in the interaction term remain free fields. If this
occurs, however, then these Standard Model modes are effectively decoupled
from the probe anyway.}
Finally, there is also a symmetry generator $T_{3}$ associated with the
tilting of the seven-brane configuration. This generator is given as a linear
combination of generators in the Cartan subalgebra of the parent group
$G_{parent}$. The precise definition is as follows \cite{FCFT}. Given a nilpotent Jordan
block decomposition of the constant part of $\Phi$:%
\begin{equation}
\Phi_{0}=\underset{a=1}{\overset{k}{\oplus}}J^{(a)},
\end{equation}
each block has an associated spin $j_{(a)}=\frac{1}{2}(n_{(a)}-1)$
representation of $SU(2)$. Denote by $T_{3}^{(a)}$ the $L_{z}$ generator of
this representation. The generator $T_{3}$ is then given as a sum of these
generators:%
\begin{equation}
T_{3}\equiv\sum_{a}T_{3}^{(a)}.
\end{equation}
In line (\ref{tabtab}) we have denoted the $T_{3}$ charge of an operator
$\mathcal{O}_{s}$ by $+s$, and that of a Standard Model field $s_{\Psi}$. Let
us note that here, we are working with respect to a particular holomorphic gauge in which
the SM field has a well-defined $T_{3}$ charge \cite{TBRANES}. See Appendices A and B for further
discussion on this point.
In terms of these symmetries, the infrared R-symmetry can be determined, much
as in \cite{FCFT}, to be a linear combination of the form:%
\begin{equation}
R_{IR}=R_{UV}+\left( \frac{t}{2}-\frac{1}{3}\right) J_{\mathcal{N}=2}%
-tT_{3}+u_{1}U_{1}+u_{2}U_{2}+\underset{\Psi}{\sum
}u_{\Psi}U_{\Psi}.
\end{equation}
Here the sum over $\Psi$ extends over all MSSM chiral superfields which
couple to the probe brane sector.
The $u_{i}$ coefficients are fixed by the condition that some of the
deformations in $\delta W_{tilt}$ are marginal in the infrared. For example, the
coefficients $u_{1}$ and $u_{2}$ are:
\begin{equation}
u_{1} =\left( S_{1}+\frac{3}{2}\right) t-1 \,\,,\,\, u_{2} =\left( S_{2}+\frac{3}{2}\right) t-1
\end{equation}
where $S_{1}$ (resp. $S_{2}$) denotes the contribution from the
deformation $\Phi(Z_{1},Z_{2})$ which is linear in $Z_{1}$ (resp. $Z_{2}$),
and has the lowest $T_{3}$ charge. We refer to the coefficients appearing in the above as:
\begin{align}
\mu_{1} =\left( S_{1}+\frac{3}{2}\right) \,\,,\,\, \mu_{2} =\left( S_{2}+\frac{3}{2}\right).
\end{align}
Similar considerations allow us to fix the $u_{\Psi}$ coefficients.
The basic condition is that if we demand that the operator
$\Psi_{R}\cdot\mathcal{O}_{\overline{R}}$ is marginal in the infrared, then we
must require $R_{IR}(\Psi_{R}^{SM}\cdot\mathcal{O}_{\overline{R}})=2$. In
holomorphic gauge, $\Psi_{R}$ and $\mathcal{O}_{\overline{R}}$ have opposite
$T_{3}$ charges \cite{TBRANES}. This imposes the condition:%
\begin{equation}\label{uPSI}
u_{\Psi}=t-\frac{2}{3}.
\end{equation}
To summarize, we introduce a net $U(1)_{SM}$ given by the sum of all
$U(1)_{\Psi}$'s which are not free fields in the infrared:%
\begin{equation}
U\left( 1\right) _{SM}=\underset{\Psi\text{ not free}}%
{\sum}U(1)_{\Psi}.
\end{equation}
The infrared R-symmetry is then:%
\begin{equation}
R_{IR}=R_{UV}+\left( \frac{t}{2}-\frac{1}{3}\right) J_{\mathcal{N}=2}%
-tT_{3}+u_{1}U_{1}+u_{2}U_{2}+\left( t-\frac{2}{3}\right) U_{SM}.
\label{ircharges}
\end{equation}
Thus, the infrared R-symmetry is determined up to a single parameter $t$ which
is fixed by $a$-maximization.
To perform $a$-maximization, we need to evaluate the cubic and
linear anomalies in $R_{IR}$. The computation is quite similar to the one
presented in \cite{FCFT}. The main idea is that although it is difficult to
compute the cubic and linear anomalies $R_{IR}^{3}$ in the IR, since
these symmetry currents can be seen in the UV, we can via anomaly matching
determine the form of these expressions in the UV as well. In the UV, however,
note that the degrees of freedom of the Standard\ Model and the probe sector
have decoupled. Hence, the cubic and linear anomalies are given by:%
\begin{align}
R_{IR}^{3} & =\, {\rm Tr}_{D3}R_{IR}^{3}+\, {\rm Tr}_{SM}R_{IR}^{3}.\\
R_{IR} & =\, {\rm Tr}_{D3}R_{IR}+\, {\rm Tr}_{SM}R_{IR}.
\end{align}
Here, $\, {\rm Tr}_{D3}$ refers to evaluating the corresponding anomaly by tracing over
just the degrees of freedom given by the UV $\mathcal{N}=2$ D3-brane probe
theory. The second contribution refers to the trace over those Standard Model
degrees of freedom which mix non-trivially with the probe in the infrared. The
key point is that because these two systems are decoupled in the UV, we can
evaluate these contributions separately.
Consider first the trace over the UV D3-brane degrees of freedom. In this
UV\ theory, there are no states charged under $U(1)_{SM}$. In other words, the
evaluation of the cubic anomaly as a function of $t$ is identical to that
already given in \cite{FCFT}. The resulting expressions are:%
\begin{align}
\, {\rm Tr}_{D3}R_{IR}^{3} & =\left( 12a_{E_{8}}-9c_{E_{8}}-\frac{3k_{E_{8}}r}{4}\right)
t^{3}+\left( -24a_{E_{8}}+12c_{E_{8}}+\frac{3u_{1}}{4}+\frac{3u_{2}}{4}\right)
t^{2}\\
& +\left( 24a_{E_{8}}-12c_{E_{8}}-\frac{3u_{1}^{2}}{2}-\frac{3u_{2}^{2}}%
{2}\right) t+(u_{1}^{3}+u_{2}^{3})\\
\, {\rm Tr}_{D3}R_{IR} & =(24a_{E_{8}}-24c_{E_{8}})t+(u_{1}+u_{2}).
\end{align}
Here, $r$ is a group theory parameter which measures the size of the Jordan block structure associated with $\Phi(0,0)$:
\begin{equation}
r = 2 \, {\rm Tr}(T_{3}T_{3}).
\end{equation}
The central charges $a_{E_{8}}$, $c_{E_{8}}$ and $k_{E_{8}}$ are the anomaly
coefficients associated with the $\mathcal{N}=2$ SCFT in the UV. For a probe
with $N$ D3-branes, the resulting values are
\cite{Cheung:1997id,Aharony:2007dj}:
\begin{align}
a_{E_{n}} & =\frac{1}{4}N^{2}\Delta+\frac{1}{2}N(\Delta-1),\\
c_{E_{n}} & =\frac{1}{4}N^{2}\Delta+\frac{3}{4}N(\Delta-1),\\
k_{E_{n}} & =2N\Delta.
\end{align}
where $N$ is the number of coincident probe D3-branes and $\Delta=6,4,3$ is
the scaling dimension of the Coulomb branch parameter for the $E_{8}$, $E_{7}$, and
$E_{6}$ Minahan-Nemeschansky theories, respectively. Here we have included the
contribution from the decoupled hypermultiplet $Z_1 \oplus Z_2$, which is why our expression is
different from the one in \cite{Aharony:2007dj}. In some cases such as where
only $Z_{1}$ couples to the configuration, there is an additional factor which
must be subtracted.
The contribution from the MSSM degrees of freedom can also be evaluated in
the UV:
\begin{align}
\, {\rm Tr}_{SM}R_{IR}^{3} & =\underset{\Psi\text{ not free}}{\sum}d_{\Psi}%
\times\left( -\frac{1}{3}-t\cdot T_{3}(\Psi)+u_{\Psi}\right) ^{3}\\
\, {\rm Tr}_{SM}R_{IR} & =\underset{\Psi\text{ not free}}{\sum}d_{\Psi}\times\left(
-\frac{1}{3}-t\cdot T_{3}(\Psi)+u_{\Psi}\right)
\end{align}
where $d_{\Psi}$ is the dimension of the representation for $\Psi$.
Putting this together, we can write the trial central charge $a_{trial}$ as:%
\begin{equation}
a_{trial}=\frac{3}{32}\left( 3\, {\rm Tr}_{D3}R_{IR}^{3}+3\, {\rm Tr}_{SM}R_{IR}^{3}%
-\, {\rm Tr}_{D3}R_{IR}-\, {\rm Tr}_{SM}R_{IR}\right) .
\end{equation}
$a$-maximization then implies that the local maximum of $a_{trial}$ as a
function of $t$ yields the value of $t=t_{\ast}$ corresponding to the infrared superconformal R-symmetry.
One consequence of this analysis is that we typically find that the Standard
Model fields develop only small anomalous dimensions. This is important because
it means that the Standard Model fields basically retain
their identity, even in the infrared theory. As we show later, this also means that in
more realistic situations, the actual suppression of the MSSM superpotential
will not be that significant. Assuming flavor is generated at the GUT scale, this
allows us to lower the CFT breaking scale below the GUT scale.\footnote{Let us note that it is in principle possible to
consider models of flavor physics where hierarchical mass patterns are generated by non-zero anomalous
dimensions \cite{Nelson:2000sn,Nelson:2001mq}. This is also a logical possibility in the present class of models. Here,
the idea would be that what is referred to as a ``third generation field'' at the GUT scale develops conformally suppressed Yukawas. Since what is
referred as the ``second generation field'' at the GUT scale couples only weakly to the probe D3-brane, this mode would, at lower energies,
become the effective third generation. Note that this leads to some suppression in the overall Yukawa matrices, unless additional fine-tuning
is included. Though it would be interesting to study such possibilities further, in this work we mainly consider the most straightforward option
that what is identified as the third generation at the GUT scale remains so at lower energies as well.}
As a final remark, let us note that there is a further class of deformations one can consider adding to the CFT,
given by allowing some of the Standard Model fields to develop non-zero vevs. For example, a non-zero Higgs vev
can trigger CFT breaking for the D3-brane sector. In this case, one can see that the resulting theory does not flow
to an interacting fixed point. This is analogous to what happens in SQCD when enough flavors get a mass to push
the theory out of the conformal window. Indeed, returning to our discussion
of the infrared R-symmetry in equation (\ref{ircharges}), we see that if we demand the deformation
$\langle \Psi_{R} \rangle \cdot \mathcal{O}_{\overline{R}}$ has R-charge $+2$ in the infrared, then the
parameter $t$ satisfies the condition $t(s + 1) = 0$ where $s$ is the $T_{3}$ charge of $\mathcal{O}_{\overline{R}}$. On
the other hand, in all T-brane examples, localized matter fields have $s < 0$, so the $\mathcal{O}$'s that they can pair
with have $s > 0$. This means that $t = 0$, which is clearly problematic for an interacting fixed point. This is
an indication that the theory is no longer conformal and instead develops a characteristic mass scale on the
order of $\langle \Psi_{R} \rangle$.
\subsection{Beta Functions}
Assuming that we do flow to an interacting CFT, we would like to establish
certain properties about how it affects the visible sector. To this end, we
now discuss how the probe sector affects the running of the visible sector
gauge couplings.\ Under the assumption (soon to be verified in a variety of
examples) that the Standard Model is only perturbed slightly by coupling to
the CFT, we can compute at weak gauge coupling the effects of an additional
threshold correction from the new states charged under the Standard
Model gauge group. From the perspective of the CFT, this amounts to computing
the two-point function for the flavor symmetry currents of $SU(3)\times
SU(2)\times U(1)$. Let us denote these currents by $J_{G}$ for $G=SU(3)$,
$SU(2)$, $U(1)$. Upon weakly gauging this current,
the one loop beta function will then be given by :%
\begin{equation}
\beta_{G}\equiv\frac{\partial g_{G}}{\partial\ln\mu}=\frac{g_{G}^{3}}%
{16\pi^{2}}b_{G},\quad\mathrm{where}\quad b_{G}=-3\,\mathrm{Tr}(R_{IR}%
J_{G}J_{G}). \label{betafunc}%
\end{equation}
Here, the \textquotedblleft Tr\textquotedblright\ trace refers to the anomaly
coefficient associated with one $R_{IR}$ current and two $J_{G}$ global
symmetry currents. In a weakly coupled setting, the trace would be over the
elementary degrees of freedom of the theory.\ This formula comes from treating
the CFT as matter interacting with the weakly gauged flavor
symmetry\footnote{For a detailed discussion, see \cite{Benini:2009mz}.}.
As a first warmup case, we can consider the contribution to the running of the $SU(5)_{GUT}$ coupling
in the limit where there are no probe/MSSM F-term couplings. This corresponds to the case where only $\delta W_{tilt}$ enters
as a deformation of the probe theory. This limiting case has been studied in \cite{FCFT}. Letting $t_{\Phi}$ denote the value of
the parameter $t$ obtained by performing $a$-maximization in this simplified case, the universal GUT contribution
from the probe D3-brane is:
\begin{equation}\label{deltabsufiv}
\delta b_{SU(5)} \equiv \frac{3 k_{E_{8}} t_{\Phi}}{4}.
\end{equation}
We shall encounter a similar contribution later for each Standard Model gauge group factor. However, to emphasize the fact that this is
associated with a GUT scale threshold, we reserve the notation $\delta b_{SU(5)}$ for equation (\ref{deltabsufiv}).
We now compute the beta function for the Standard Model gauge group
factors. Near the scale where the MSSM superpartners first enter the
spectrum, the corresponding one-loop MSSM\ beta functions are, in our sign conventions:%
\begin{equation}
b_{SU(3)}^{(0)}=-3\text{, }b_{SU(2)}^{(0)}=+1\text{, }b_{U(1)}^{(0)}%
=+\frac{33}{5}.
\label{origbeta}
\end{equation}
In a step function approximation to the running, this is the value of the beta
function until the additional states from the probe sector enter the spectrum.
At this point, additional contributions will affect the running.
Turning the discussion around, we can proceed from high energies down to
lower energies. There will then be the usual contribution from $b^{(0)}$, as
well as additional contributions induced by the presence of the probe sector.
As indicated in equation (\ref{betafunc}), the one-loop beta function is given
by an anomaly coefficient. In the IR, it is difficult to directly list all of
the degrees of freedom, because the probe and Standard Model now interact
non-trivially. Note, however, that because $b_{G}$ is an anomaly coefficient,
we can use anomaly matching to compute it in the UV, where the Standard Model
and probe sector degrees of freedom are decoupled. The key point for us is
that $R_{IR}$ is a linear combination of $R_{UV}$ and flavor symmetries of the
UV theory, and $J_{G}$ is a flavor symmetry present in both the UV\ and IR
theories.
In terms of the UV degrees of freedom, we then have:%
\begin{equation}
b_{G}=-3\,\mathrm{Tr}(R_{IR}J_{G}J_{G})=-3\,\mathrm{Tr}_{SM}(R_{IR}J_{G}%
J_{G})-3\,\mathrm{Tr}_{D3}(R_{IR}J_{G}J_{G}).
\end{equation}
Using the form of the IR R-symmetry (\ref{ircharges}) and the charge assignments in (\ref{tabtab}),
we find that the shifts in the beta functions
\begin{equation}
\delta b_{G} \equiv b_{G}-b_{G}^{(0)}%
\end{equation}
for a gauge group factor $G$ are given by:%
\begin{align}
\delta b_{SU(3)} & =\frac{3k_{E_{8}}}{4}t+\frac{9}{2}
\times\left( t\cdot T_{3}^{(10_{M})}-u_{10_{M}}\right) +\frac{3}{2}\times\left( t\cdot
T_{3}^{(\overline{5}_{M})}-u_{\overline{5}_{M}}\right) \\
\delta b_{SU(2)} & =\frac{3k_{E_{8}}}{4}t+\frac{9}{2}%
\times\left( t\cdot T_{3}^{(10_{M})}-u_{10_{M}}\right) +\frac{3}{2}\times\left( t\cdot
T_{3}^{(\overline{5}_{M})}-u_{\overline{5}_{M}}\right) \\
& \qquad\qquad\,\,+\frac{3}{2}\times\left( t\cdot T_{3}^{(H_{d})}-u_{H_{d}%
}\right) +\frac{3}{2}\times\left( t\cdot T_{3}^{(H_{u})}-u_{H_{u}}\right) \\
\delta b_{U(1)} & =\frac{3k_{E_{8}}}{4}t+\frac{9}{2}%
\times\left( t\cdot T_{3}^{(10_{M})}-u_{10_{M}}\right) +\frac{3}{2}\times\left( t\cdot
T_{3}^{(\overline{5}_{M})}-u_{\overline{5}_{M}}\right) \\
& +3\times\frac{18}{60}\times\left( t\cdot T_{3}^{(H_{d})}-u_{H_{d}}\right)
+3\times\frac{18}{60}\times\left( t\cdot T_{3}^{(H_{u})}-u_{H_{u}}\right) .
\end{align}
Here, we are assuming that all modes have developed an anomalous dimension. The corresponding terms which are
free fields are to be omitted from this expression. In all cases, the first line is the same for $SU(3)$, $SU(2)$ and $U(1)$. In particular, such contributions will not distort gauge coupling unification. The additional contributions from the Higgs fields to each gauge group factor are
different. We shall later comment on the effects of these shifts, and their (helpful)
consequences for precision unification. Note also that these shifts are missing the contribution
from $R_{UV}$ in (\ref{ircharges}), since that effect is contained in $b_G^{(0)}$ and has
been subtracted off. Finally, note that our normalization for the $U(1)$ generator is chosen to agree with a canonical
embedding in $SU(5)_{GUT}$.
\subsection{Example: $\mathbb{Z}_{2}$ Monodromy}
To illustrate the main ideas discussed earlier, in this section
we consider a simplified model based on a $\mathbb{Z}_{2}$
monodromy group which realizes a visible sector with the couplings
$5_H \times 10_M \times 10_M$ and $\overline{5}_H \times \overline{5}_M \times 10_M$. These
examples do not include a neutrino scenario as in \cite{BHSV,EPOINT}. However,
given the extra flexibility afforded by D3-branes, there may be novel ways to
include neutrinos in such setups. Since they are among the simplest examples
of T-brane configurations we consider these examples first.
The coarse-grained T-brane configuration we consider is defined by:
\begin{equation}
\Phi=\left[
\begin{array}
[c]{ccccc}%
0 & 1 & & & \\
Z_{1} & 0 & & & \\
& & 0 & & \\
& & & 0 & \\
& & & & 0
\end{array}
\right] . \label{deformar}%
\end{equation}
The characteristic polynomial for $\Phi$ has Galois group $\mathbb{Z}_{2}$, which is identified with the ``monodromy group''.
Higher order (irrelevant to the IR D3-brane theory) terms serve to fix the profile of
localized matter in the geometry. As explained in Appendix A, such
considerations are not so important for the considerations of this paper. The
$T_{3}$ generator is given by:%
\begin{equation}
T_{3}=\text{diag}(1/2,-1/2,0,0,0).
\end{equation}
In a singular branched gauge, we can describe this scenario as having a $%
\mathbb{Z}
_{2}$ monodromy group which acts on the eigenvalues $\lambda_{i}$ of $\Phi$
by permuting $\lambda_{1}$ and $\lambda_{2}$, while keeping the other $\lambda_{i}$ fixed.
In terms of the eigenvalues $\lambda_{i}$,
we have that the visible sector modes fill out the
following orbits under the $%
\mathbb{Z}
_{2}$ monodromy group:%
\begin{align}
10_{M} & :\{\lambda_{1},\lambda_{2}\}\\
5_{H} & :\{-\lambda_{1}-\lambda_{2}\}\\
\overline{5}_{1} & :\{\lambda_{1}+\lambda_{3},\lambda_{2}+\lambda_{3}\}\\
\overline{5}_{2} & :\{\lambda_{4}+\lambda_{5}\}.
\end{align}
The subscript on the $\overline{5}$ fields indicates that in this simple
example, we can interchange the roles of the $\overline{5}_{M}$ and
$\overline{5}_{H}$. We consider both possibilities in what follows, and refer to the monodromy
scenarios as $\mathbb{Z}^{(1)}_{2}$ and $\mathbb{Z}^{(2)}_{2}$.
Based on our general discussion of probe/MSSM couplings in Appendix A, we can
determine the vector transforming in a representation of $SU(5)_{\bot}$ which
dominantly couples to the D3-brane, as well as its corresponding $T_{3}$
charge:
\begin{align}
&
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $H_{u}$ & $H_{d}$ & $\overline{5}_{M}$ & $10_{M}$\\\hline
Vector & $e_{1}^{\ast}\wedge e_{2}^{\ast}$ & $e_{2}\wedge e_{3}$ &
$e_{4}\wedge e_{5}$ & $e_{2}$\\\hline
$T_{3}$ charge & $0$ & $-1/2$ & $0$ & $-1/2$\\\hline
\end{tabular}
\\
&
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $H_{u}$ & $H_{d}$ & $\overline{5}_{M}$ & $10_{M}$\\\hline
Vector & $e_{1}^{\ast}\wedge e_{2}^{\ast}$ & $e_{4}\wedge e_{5}$ &
$e_{2}\wedge e_{3}$ & $e_{2}$\\\hline
$T_{3}$ charge & $0$ & $0$ & $-1/2$ & $-1/2$\\\hline
\end{tabular}
\text{ \ \ \ }.
\end{align}
\subsubsection*{$\Phi$-deformed Theory}
To begin our analysis of the probe/MSSM fixed points, we first recall some
properties of the CFT obtained in the limit in which all Standard Model fields
are decoupled. In other words, we first study the theory obtained by the
deformation%
\begin{equation}
\delta W_{tilt} =\, {\rm Tr}_{E_{8}}\left( \Phi(Z_{1},Z_{2})\cdot\mathcal{O}\right)
\end{equation}
with $\Phi$ given as in equation (\ref{deformar}). The resulting infrared
fixed point has basically been obtained in \cite{FCFT}. The parameters
$\mu_{1}=5/2$, $r=1$. The value of $\mu_{2}$ has \ no physical meaning in this
case, because $Z_{2}$ remains decoupled in the infrared. Via $a$-maximization,
we can determine the critical value of $t$ which maximizes $a$, which we
denote by $t_{\Phi}$. Its value is $t_{\Phi} = 0.53$.
We can also compute the infrared values of the central
charges $a$, $c$ and $k$. To properly compare different theories, we also
include the contributions to $a$ and $c$ from the UV\ decoupled modes $Z_{2},$
as well as the decoupled Standard Model fields $H_{u}$, $H_{d}$, $\overline
{5}_{M}$ and $10_{M}$. Finally, we also consider the contribution to the
running of the $SU(5)_{GUT}$ coupling from just the D3-brane sector, given by
$\delta b_{SU(5)}=3k_{E_{8}}t_{\ast}/4$:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
& $t_{\Phi}$ & $a_{\Phi}$ & $c_{\Phi}$ & $\delta b_{SU(5)}$\\\hline
$%
\mathbb{Z}
_{2}$ & $0.53$ & $3.83$ & $5.21$ & $4.80$\\\hline
\end{tabular}
\end{equation}
Given the values of these parameters, we can also determine the scaling
dimensions of the operators which eventually will couple to the Standard Model
fields. This is dictated by the $T_{3}$ charge assignments, and so will depend
on the particular monodromy scenario we consider. In our case, we have
operators with $T_{3}$ charge $0$ or $+1/2$. This leads to scaling dimensions:%
\begin{align}
\Delta\left( \mathcal{O}_{s=0}\right) & =3-\frac{3}{2}t\\
\Delta\left( \mathcal{O}_{s=1/2}\right) & =3-\frac{9}{4}t
\end{align}
We can use this to determine which of our original deformations are relevant,
or marginal in the IR\ of the $\Phi$-deformed theory:
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
Dimensions & $\mathcal{O}_{H_{u}}$ & $\mathcal{O}_{H_{d}}$ & $\mathcal{O}%
_{\overline{5}_{M}}$ & $\mathcal{O}_{10_{M}}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $2.20$ & $1.80$ & $2.20$ & $1.80$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $2.20$ & $2.20$ & $1.80$ & $1.80$\\\hline
\end{tabular}
\ \ \ \ \ \ \ .
\end{equation}
\subsubsection*{Infrared Theory}
Let us now compute the form of the infrared R-symmetry in the two $\mathbb{Z}_{2}$
monodromy scenarios. Again, we must evaluate the various cubic and linear anomalies.
Here, we have:
\begin{align}
\mathbb{Z}^{(1)}_{2}\text{ Case} & \text{: }R_{IR}^{3}=\, {\rm Tr}_{D3}R_{IR}%
^{3}+(2+10)\times\left( -\frac{1}{3}+\frac{t}{2}+u_{\Psi}\right) ^{3}\\
\mathbb{Z}_{2}^{(2)}\text{ Case} & \text{: }R_{IR}^{3}=\, {\rm Tr}_{D3}R_{IR}%
^{3}+(5+10)\times\left( -\frac{1}{3}+\frac{t}{2}+u_{\Psi}\right) ^{3}%
\end{align}
where in the above $u_{\Psi} = t - 2/3$ as in equation (\ref{uPSI}).
Similarly, we can evaluate the anomaly linear in R-charge. We have:%
\begin{align}
\mathbb{Z}_{2}^{(1)}\text{ Case} & \text{: }R_{IR}=\, {\rm Tr}_{D3}R_{IR}%
+(2+10)\times\left( -\frac{1}{3}+\frac{t}{2}+u_{\Psi}\right) \\
\mathbb{Z}_{2}^{(2)}\text{ Case} & \text{: }R_{IR}=\, {\rm Tr}_{D3}R_{IR}%
+(5+10)\times\left( -\frac{1}{3}+\frac{t}{2}+u_{\Psi}\right) .
\end{align}
Performing $a$-maximization in these two cases yields the same value (after
rounding):%
\begin{equation}
t_{\ast}=0.50. \label{tmaxone}%
\end{equation}
From this, we conclude that the IR\ dimensions of the Standard Model fields,
and $\mathcal{O}$ operators are:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|c|}\hline
IR\ Dimensions & $H_{u}$ & $\mathcal{O}_{H_{u}}$ & $H_{d}$ & $\mathcal{O}%
_{H_{d}}$ & $\overline{5}_{M}$ & $\mathcal{O}_{\overline{5}_{M}}$ & $10_{M}$ &
$\mathcal{O}_{10_{M}}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $1$ & $2.25$ & $1.13$ & $1.87$ & $1$ & $2.25$ & $1.13$ &
$1.87$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $1$ & $2.25$ & $1$ & $2.25$ & $1.12$ & $1.88$ & $1.12$ &
$1.88$\\\hline
\end{tabular}
\ \ \ \ \ \ \ \ \ .
\end{equation}
Let us consider the suppression of the Yukawa couplings associated
with the superpotential couplings~$W_{MSSM}$. For the modes which have maximal
coupling to the probe D3-brane sector, we have, at the CFT breaking scale:%
\begin{equation}
\lambda_{5\times10\times10}\sim\lambda_{\overline{5}\times\overline{5}%
\times10}\sim\left( \frac{M_{\cancel{CFT}}}{M_{GUT}}\right) ^{0.26}.
\end{equation}
For $M_{\cancel{CFT}}/M_{GUT}\sim10^{-3}$, this leads to $\lambda\sim0.2$,
which is only a mild suppression. This is easily compensated for
by a small enhancement of the Yukawa at the GUT scale.
We can also compute the values of the central charges $a$ and $c$, and compare
with their values in the original $\mathcal{N}=2$ and $\Phi$-deformed
theories. Here, the UV theory contains both the contributions to $a$ and $c$
from the $\mathcal{N}=2$ probe D3-brane, as well as the decoupled
Standard\ Model fields and the decoupled hypermultiplet $Z_{1}\oplus Z_{2}$.
We find:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|c|c|}\hline
Central Charges & $a_{UV}$ & $a_{\Phi}$ & $a_{IR}$ & $c_{UV}$ & $c_{\Phi}$ &
$c_{IR}$\\\hline
$\mathbb{Z}_{2}^{(1)}$ & $4.40$ & $3.83$ & $3.79$ & $6.04$ & $5.21$ &
$5.06$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $4.40$ & $3.83$ & $3.79$ & $6.04$ & $5.21$ & $5.03$\\\hline
\end{tabular}
\ \ \ \ \ \ \ \ \ .
\end{equation}
Note that in both scenarios, there is a decrease in the values of the central
charges. The rather mild decrease provides further evidence that there is
only weak mixing between the probe and Standard Model.
Let us now compute the contribution to the SM beta functions.
Basically, this is dictated by the $T_{3}$ charge
assignments of those visible sector fields which have
developed a non-trivial scaling dimension in the IR theory. First consider the
$%
\mathbb{Z}
_{2}^{(1)}$ scenario. Here, we have:
\begin{align}
\delta b_{SU(3)} & =\frac{3k_{E_{8}}}{4}t-\frac{9}{2}\times\left( \frac
{t}{2}+u_{10_{M}}\right) \\
\delta b_{SU(2)} & =\frac{3k_{E_{8}}}{4}t-\frac{9}{2}\times\left( \frac
{t}{2}+u_{10_{M}}\right) \,-\frac{3}{2}\times\left( \frac{t}{2}+u_{H_{d}%
}\right) \\
\delta b_{U(1)} & =\frac{3k_{E_{8}}}{4}t-\frac{9}{2}\times\left( \frac
{t}{2}+u_{10_{M}}\right) -3\times\frac{18}{60}\left( \frac{t}{2}+u_{H_{d}%
}\right) .
\end{align}
Next consider the $%
\mathbb{Z}
_{2}^{(2)}$ scenario. Here, we have:%
\begin{align}
\delta b_{SU(3)} & =\frac{3k_{E_{8}}}{4}t -\frac{9}{2}\times\left( \frac{t}%
{2}+u_{10_{M}}\right) -\frac{3}{2} \times \left( \frac{t}%
{2}+u_{\overline{5}_{M}}\right) \\
\delta b_{SU(2)} & =\frac{3k_{E_{8}}}{4}t -\frac{9}{2}\times\left( \frac{t}%
{2}+u_{10_{M}}\right) -\frac{3}{2}\times\left( \frac
{t}{2}+u_{\overline{5}_{M}}\right) \,\\
\delta b_{U(1)} & =\frac{3k_{E_{8}}}{4}t -\frac{9}{2}\times\left( \frac{t}%
{2}+u_{10_{M}}\right) -\frac{3}{2}\times\left( \frac
{t}{2}+u_{\overline{5}_{M}}\right)
\end{align}
which is a universal shift to the one-loop MSSM\ beta functions. Using the
expression $u=t-2/3$ and plugging in the value of $t_{\ast} \sim0.50$ yields:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|}\hline
Beta Functions & $\delta b_{SU(3)}$ & $\delta b_{SU(2)}$ & $\delta b_{U(1)}%
$\\\hline
\multicolumn{1}{|c|}{$%
\mathbb{Z}
_{2}^{(1)}$} & $4.13$ & $4.$ & $4.05$\\\hline
\multicolumn{1}{|c|}{$%
\mathbb{Z}
_{2}^{(2)}$} & $4.$ & $4.$ & $4.$\\\hline
\end{tabular}
\ \ \ \ \ \ \ \ .
\end{equation}
\subsection{Summary of Fixed Points \label{sec:opbetafunc}}
Repeating a similar analysis as above for different monodromy choices, we can
find the anomalous dimensions for CFT and MSSM fields. From the perspective of
the CFT, the main thing we must specify is the $T_{3}$ charge of the various
operators. Below we summarize the operators and their corresponding $T_{3}$
charges, as well as the group-theoretic parameters $\mu_{1},\mu_{2}$ and $r$
entering into $R_{IR}$:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|}\hline
$R_{IR}$\ Parameters & $\mu_{1}$ & $\mu_{2}$ & $r$ & $T_{3}(H_{u})$ &
$T_{3}\left( H_{d}\right) $ & $T_{3}\left( \overline{5}_{M}\right) $ &
$T_{3}\left( 10_{M}\right) $\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $5/2$ & X & $1$ & $0$ & $-1/2$ & $0$ & $-1/2$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $5/2$ & X & $1$ & $0$ & $0$ & $-1/2$ & $-1/2$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $5/2$ & $5/2$ & $2$ & $0$ & $-1$ & $-1/2$ & $-1/2$\\\hline
$S_{3}$ & $7/2$ & $5/2$ & $4$ & $-1$ & $-1$ & $-1$ & $-1$\\\hline
$Dih_{4}^{(1)}$ & $9/2$ & $5/2$ & $10$ & $-2$ & $-3/2$ & $0$ & $-3/2$\\\hline
$Dih_{4}^{(2)}$ & $9/2$ & $5/2$ & $10$ & $0$ & $-3/2$ & $-2$ & $-3/2$\\\hline
\end{tabular}
\ \ \ \ \text{ \ \ \ \ }.
\end{equation}
An ``X'' indicates the entry of the table has no meaning.
See Appendix B for the definition of each monodromy scenario, and further
explanation of how the $T_{3}$ charge assignments are fixed for the modes
which dominantly couple to the probe D3-brane.
Given these values of the parameters, we can determine the form of the
infrared R-symmetry. For each scenario, those Standard Model fields which have
negative $T_{3}$ charge are those which are charged under $U(1)_{SM}$. Hence,
we can compute the cubic and linear anomalies, much as we did abstractly in
the previous sections, as well as in the explicit $%
\mathbb{Z}
_{2}$ monodromy scenarios described explicitly in the previous subsection.
Since the computation is rather similar in all cases, we shall omit the
details. Performing $a$-maximization, we can then extract various properties of
the IR theories. In all cases, the rather small shift in going from the
critical value of $t_{\Phi}$ of the intermediate $\Phi$-deformed theories,
to the eventual value of $t_{\ast}$ in the coupled probe/MSSM fixed point means that just as in
\cite{FCFT}, there are no obvious unitarity bound violations. We now turn to
the one and two D3-brane probe scenarios.
\subsubsection{Scenarios with One D3-Brane}
In this subsection we summarize the main computable properties of a single D3-brane probing a
particular configuration of seven-brane monodromy. To this end, we list the IR scaling dimensions of the
Higgs fields and third generation chiral matter fields, e.g. those which dominantly couple to the probe. We also
include the scaling dimensions of the corresponding $\mathcal{O}$ operators:
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|c|c|}\hline
IR Dimensions & $t_{\ast}$ & $H_{u}$ & $\mathcal{O}_{H_{u}}$ & $H_{d}$ &
$\mathcal{O}_{H_{d}}$ & $\overline{5}_{M}$ & $\mathcal{O}_{\overline{5}_{M}}$
& $10_{M}$ & $\mathcal{O}_{10_{M}}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $0.50$ & $1$ & $2.25$ & $1.13$ & $1.87$ & $1$ & $2.25$ & $1.13$
& $1.87$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $0.50$ & $1$ & $2.25$ & $1$ & $2.25$ & $1.12$ & $1.88$ & $1.12$
& $1.88$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $0.45$ & $1$ & $2.33$ & $1.35$ & $1.65$ & $1.01$ & $1.99$ & $1.01$ &
$1.99$\\\hline
$S_{3}$ & $0.36$ & $1.08$ & $1.92$ & $1.08$ & $1.92$ & $1.08$ & $1.92$ &
$1.08$ & $1.92$\\\hline
$Dih_{4}^{(1)}$ & $0.28$ & $1.25$ & $1.75$ & $1.04$ & $1.96$ & $1$ & $2.58$ &
$1.04$ & $1.96$\\\hline
$Dih_{4}^{(2)}$ & $0.27$ & $1$ & $2.59$ & $1.02$ & $1.98$ & $1.22$ & $1.78$ &
$1.02$ & $1.98$\\\hline
\end{tabular}
\end{equation}
In nearly all cases, we observe that there is only a
small shift in the scaling dimension of the Standard Model fields.
Additionally, we list the central charges $a$ and $c$ for the various
scenarios. To properly compare various theories, we include the contributions from $Z_1$, $Z_2$, $H_u$, $H_d$, $\overline{5}_{M}$ and $10_{M}$,
even if these modes are decoupled in the IR. Along these lines, we have included in $a_{UV}$ the contribution from the $\mathcal{N} = 2$ probe D3-brane,
as well as the decoupled chiral multiplets. Similar considerations apply for $c_{UV}$, $a_{\Phi}$ and $c_{\Phi}$:
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|c|c|}\hline
Central Charges & $a_{UV}$ & $a_{\Phi}$ & $a_{IR}$ & $c_{UV}$ & $c_{\Phi}$ &
$c_{IR}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $4.40$ & $3.83$ & $3.79$ & $6.04$ & $5.21$ & $5.06$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $4.40$ & $3.83$ & $3.79$ & $6.04$ & $5.21$ & $5.03$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $4.40$ & $3.50$ & $3.48$ & $6.04$ & $4.74$ & $4.66$\\\hline
$S_{3}$ & $4.40$ & $3.08$ & $3.05$ & $6.04$ & $4.19$ & $4.05$\\\hline
$Dih_{4}^{(1)}$ & $4.40$ & $2.49$ & $2.47$ & $6.04$ & $3.43$ & $3.35$\\\hline
$Dih_{4}^{(2)}$ & $4.40$ & $2.49$ & $2.45$ & $6.04$ & $3.43$ & $3.31$\\\hline
\end{tabular}
\end{equation}
In the above collection of numbers, the parameters $a_{\Phi}$ and $c_{\Phi}$
refer to \textquotedblleft intermediate\textquotedblright\ values of the
central charges given by the IR\ theory with all couplings to the Standard
Model switched off.
Finally, we also list the contributions to the beta functions for
the various scenarios. Here, the parameter $\delta b_{SU(5)}$ refers
to the contribution to the running of $SU(5)_{GUT}$ in the limit where the
F-term couplings to the Standard Model have been switched off, just as in equation (\ref{deltabsufiv}).
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
Beta Functions & $\delta b_{SU(5)}$ & $\delta b_{SU(3)}$ & $\delta b_{SU(2)}$
& $\delta b_{U(1)}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $4.80$ & $4.13$ & $4.$ & $4.05$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $4.80$ & $4.$ & $4.$ & $4.$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $4.16$ & $4.00$ & $3.65$ & $3.79$\\\hline
$S_{3}$ & $3.50$ & $2.92$ & $2.75$ & $2.81$\\\hline
$Dih_{4}^{(1)}$ & $2.63$ & $2.38$ & $2.08$ & $2.20$\\\hline
$Dih_{4}^{(2)}$ & $2.63$ & $2.16$ & $2.14$ & $2.15$\\\hline
\end{tabular}
\ \ \ \ \ \text{ \ \ \ \ }.
\end{equation}
Note that the values of the beta functions are all consistent with the
requirement that $SU(5)_{GUT}$ remains asymptotically free. Further
note that in passing from the $\Phi$-deformed theory to the coupled probe/MSSM
system, there is only a small change in both the central charges, and the beta
functions. This is consistent with the expectation that the probe and Standard
Model only weakly mix.
\subsubsection{Scenarios with Two D3-Branes}
We can perform a similar analysis for the probe theories involving two
D3-branes. Our conventions are the same as in the case of a single D3-brane.
Basically, the only change is to the parameter $N$ entering into
the $\mathcal{N} = 2$ central charges $a_{E_{8}}$, $c_{E_{8}}$ and $k_{E_{8}}$.
The IR scaling dimensions are:
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|c|c|}\hline
IR Dimensions & $t_{\ast}$ & $H_{u}$ & $\mathcal{O}_{H_{u}}$ & $H_{d}$ &
$\mathcal{O}_{H_{d}}$ & $\overline{5}_{M}$ & $\mathcal{O}_{\overline{5}_{M}}$
& $10_{M}$ & $\mathcal{O}_{10_{M}}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $0.54$ & $1$ & $2.19$ & $1.22$ & $1.78$ & $1$ & $2.19$ & $1.22$
& $1.78$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $0.54$ & $1$ & $2.19$ & $1$ & $2.19$ & $1.21$ & $1.79$ & $1.21$
& $1.79$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $0.48$ & $1$ & $2.27$ & $1.45$ & $1.55$ & $1.09$ & $1.91$ & $1.09$ &
$1.91$\\\hline
$S_{3}$ & $0.40$ & $1.21$ & $1.79$ & $1.21$ & $1.79$ & $1.21$ & $1.79$ &
$1.21$ & $1.79$\\\hline
$Dih_{4}^{(1)}$ & $0.31$ & $1.41$ & $1.59$ & $1.17$ & $1.83$ & $1$ & $2.53$ &
$1.17$ & $1.83$\\\hline
$Dih_{4}^{(2)}$ & $0.31$ & $1$ & $2.54$ & $1.16$ & $1.84$ & $1.39$ & $1.61$ &
$1.16$ & $1.84$\\\hline
\end{tabular}
\end{equation}
We observe that these scaling dimensions are slightly larger than their single D3-brane counterparts. However,
the overall size is still on the small side.
Next consider the central charges of the UV, intermediate, and IR theories:
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|c|c|}\hline
Central Charges & $a_{UV}$ & $a_{\Phi}$ & $a_{IR}$ & $c_{UV}$ & $c_{\Phi}$ &
$c_{IR}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $11.4$ & $10.2$ & $10.1$ & $14.3$ & $12.7$ & $12.4$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $11.4$ & $10.2$ & $10.1$ & $14.3$ & $12.7$ & $12.4$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $11.4$ & $9.45$ & $9.39$ & $14.3$ & $11.7$ & $11.5$\\\hline
$S_{3}$ & $11.4$ & $8.43$ & $8.31$ & $14.3$ & $10.4$ & $10.0$\\\hline
$Dih_{4}^{(1)}$ & $11.4$ & $6.85$ & $6.77$ & $14.3$ & $8.46$ & $8.20$\\\hline
$Dih_{4}^{(2)}$ & $11.4$ & $6.85$ & $6.72$ & $14.3$ & $8.46$ & $8.11$\\\hline
\end{tabular}
\end{equation}
As expected, the central charge for the two D3-brane system is larger than that of the single D3-brane case. Note, however,
that the decrease in going from the $\Phi$-deformed theory to the coupled probe/MSSM system is still small.
Finally, we can also compute the values of the beta functions for these scenarios:
\begin{equation}
\begin{tabular}
[c]{|c|c|c|c|c|}\hline
Beta Functions & $\delta b_{SU(5)}$ & $\delta b_{SU(3)}$ & $\delta b_{SU(2)}$
& $\delta b_{U(1)}$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $10.1$ & $9.10$ & $8.88$ & $8.96$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $10.1$ & $8.84$ & $8.84$ & $8.84$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $8.96$ & $8.36$ & $7.91$ & $8.09$\\\hline
$S_{3}$ & $7.70$ & $6.42$ & $6.$ & $6.17$\\\hline
$Dih_{4}^{(1)}$ & $5.90$ & $5.11$ & $4.53$ & $4.76$\\\hline
$Dih_{4}^{(2)}$ & $5.90$ & $4.69$ & $4.54$ & $4.60$\\\hline
\end{tabular}
\end{equation}
The values of the beta functions in most scenarios are significantly higher
than their single D3-brane counterparts. This leads to accelerated running of
the gauge couplings and typically a loss of asymptotic freedom for
$SU(5)_{GUT}$. Note, however, that in the \textquotedblleft large monodromy group
scenarios\textquotedblright\ such as the $S_{3}$ and the $Dih_{4}$ examples, the
effect is somewhat milder. A similar analysis can be performed for more than two
D3-branes. In all T-brane scenarios considered here, we lose asymptotic freedom for
$SU(5)_{GUT}$, so we do not entertain this possibility further.
\section{Threshold Corrections and Unification\label{sec:UNIFY}}
In the previous section we determined various properties of the fixed point
associated with coupling the MSSM to a strongly coupled CFT, and rather
surprisingly found that the effects on the Standard Model are mild. We have
also seen that coupling to the probe sector influences the running of the
gauge couplings, but in a way which is not $SU(5)_{GUT}$ universal. Indeed,
though the probe sector states fill out complete GUT multiplets, their
couplings to the Higgs sector explicitly break $SU(5)_{GUT}$. In this section
we discuss the consequences of such couplings for precision unification.
A compelling motivation for supersymmetric GUT theories is that at one loop order,
the gauge couplings of the MSSM appear to unify at a scale of order $\sim 10^{16}$ GeV
\cite{Dimopoulos:1981yj, Dimopoulos:1981zb, Ibanez:1981yh, Einhorn:1981sx, Marciano:1981un}. Beyond the
one-loop approximation, however, various effects can potentially distort unification. These distortions arise both
from low energy effects associated with two-loop contributions involving just the
MSSM\ degrees of freedom, as well as effects closer to the GUT scale.
Including only two-loop effects from the MSSM, it is
well-known that for typical superpartner masses,
if one runs the observed values of the gauge couplings to
higher energy scales without including any other threshold corrections, the
gauge coupling constants no longer unify. Defining $M_{GUT}$ as the energy
scale at which $\alpha_{1}(M_{GUT})=\alpha_{2}(M_{GUT})=\alpha_{GUT}$,
the value of $\alpha_{3}(M_{GUT})$ is lower than $\alpha_{GUT}$, and differs
from it by an order $4\%$ amount:%
\begin{equation}
\text{Two-Loop MSSM\ contribution: }\frac{\alpha_{3}^{-1}(M_{GUT}%
)-\alpha_{GUT}^{-1}}{\alpha_{GUT}^{-1}}\sim+4\%.
\end{equation}
See \cite{Raby:2006sk} for a recent review of such issues.
The precise amount of mismatch depends on the details of the superpartner mass
spectrum (see for example \cite{Langacker:1995fk, Raby:2009sf}). It is
equally well-known that various GUT scale threshold corrections from incomplete
GUT multiplets can induce an appropriate shift which can eliminate this
discrepancy (see for example
\cite{Ibanez:1991zv,Ibanez:1992hc,Nilles:1995kb,Hall:2001pg}). On the one
hand, this provides a way to retain gauge coupling unification. However, this
unification is achieved at the expense of including extra incomplete multiplets beyond the
Higgs fields. Let us note that such multiplets certainly exist in higher
dimensional theories such as F-theory GUTs, and are associated with the
Kaluza-Klein spectrum of excitations for the Higgs doublets and triplets.
Gauge coupling unification in F-theory GUTs has been studied in for example
\cite{DWII, Blumenhagen:2008aw, Conlon:2009qa, Marsano:2009wr, Leontaris:2009wi, Li:2010dp, Leontaris:2011pu}.
In the specific context of F-theory GUTs, a common way to break the GUT group
is through the introduction of hyperflux \cite{BHVII,DWII}. This introduces a
further distortion of gauge coupling unification, which is of the same order
and sign as the \textquotedblleft two-loop discrepancy\textquotedblright%
\ from purely MSSM\ effects \cite{DWII, Blumenhagen:2008aw}. For appropriate
values of the mass spectrum for the Higgs doublets and triplets, precision
unification can be retained \cite{DWII}. Though adding various thresholds from
incomplete GUT multiplets provides a potential way to rectify precision
unification, it is aesthetically displeasing that the least unified
parts of a Grand Unified Theory would somehow be destined to play the role of ensuring gauge
coupling unification.
In this section we show that the extra couplings between the probe sector
fields and the MSSM matter fields induce a correction to the running of the
gauge couplings which can counter these deleterious shifts to gauge
coupling unification. Moreover, this is achieved by adding a vector-like
sector with states which fill out complete GUT multiplets. The basic point,
however, is that because of the coupling to the Higgs sector, a shift is
induced in the running of the couplings.
At a qualitative level, the reason such couplings help with unification is due
to the ways that various distortions of gauge coupling unification enter. To
this end, it is helpful to recall that the numerator of the NSVZ beta function \cite{Novikov:1983uc} for a
gauge group $G$ is, in our sign conventions:
\begin{equation}
b_{G}^{NSVZ}=-3C_{2}(G)+\sum_{\Psi}C_{2}(R_{\Psi})(1-\gamma_{\Psi}).
\end{equation}
where $C_{2}(G)$ and $C_{2}(R_{\Psi})$ are the Dynkin indices
for the adjoint and representation $R_{\Psi}$, respectively. Further,
$\gamma_{\Psi}$ is the anomalous dimension associated with an elementary
field $\Psi$, and the sum is over the various elementary degrees of freedom of
the system.\footnote{In a strongly coupled non-Lagrangian theory there are
various subtleties associated with defining the elementary degrees of freedom.
In this paper we have bypassed this point by appealing to anomaly matching
considerations.} The main point is that the anomalous dimensions $\gamma$ of
the system also affect the running of the visible sector couplings. Indeed, the
two-loop distortion from MSSM\ effects comes from gauge
interactions. This contribution can be counteracted by
the effects of F-term couplings involving the Higgs fields. Such F-term couplings tend to
increase the anomalous dimensions of fields, which in turn modifies the form of the beta function.
As reviewed for example in \cite{Donkin:2010ta}, the top quark Yukawa is by itself not large enough to
correct the two-loop discrepancy generated by gauge interaction effects.
However, F-term couplings to additional sectors can lead to further shifts
which can counteract the two-loop MSSM\ discrepancy \cite{Donkin:2010ta}%
.\footnote{In \cite{Donkin:2010ta} it was assumed that the Higgs fields couple
to additional hidden sector fields through cubic terms involving two
additional matter fields charged under the Standard Model gauge group.
However, the qualitative effect is more general, and just requires the
existence of extra Yukawa couplings to the Higgs fields.} This is precisely
the situation we are in.
Here, we use our analysis of beta functions performed in the previous section
to estimate the size of threshold corrections to gauge coupling unification
induced by the probe D3-brane. Because there are various contributions to the
values of the GUT\ scale couplings, such as those induced by GUT\ scale
thresholds, e.g. hyperflux contributions and Kalua-Klein\ Higgs doublets and
triplets, as well as two-loop MSSM effects, we shall characterize all of these
effects in terms of a finite shift to the GUT\ scale values of the
inverse fine structure constants. From this perspective, precision unification
is achieved when the contribution from the one loop MSSM\ beta function and
the probe D3-brane threshold is:%
\begin{equation}
\text{Probe D3-brane\ contribution: }\frac{\alpha_{3}^{-1}(M_{GUT}%
)-\alpha_{GUT}^{-1}}{\alpha_{GUT}^{-1}}\sim- 4\%.
\end{equation}
Let us now turn to a more detailed discussion of threshold effects from the
probe sector. At some threshold scale $M_{thresh}\sim M_{\cancel{CFT}}$ we
assume that the additional CFT states enter. In a step function
approximation, the running of the various gauge couplings are:%
\begin{align}
\alpha_{3}^{-1}(l) & =\left(- \frac{b_{SU(3)}^{(0)}}{2\pi}\cdot
(l-l_{0})+\alpha_{3}^{-1}(l_{0})\right) \times\theta\left( l_{t}-l\right)
\\
& +\left(- \frac{b_{SU(3)}^{(0)} + \delta b_{SU(3)}}{2\pi}\cdot(l-l_{t})-\frac
{b_{SU(3)}^{(0)}}{2\pi}\cdot(l_{t} - l_{0})+\alpha_{3}^{-1}(l_{0})\right)
\times\theta\left( l-l_{t}\right) \\
\alpha_{2}^{-1}(l) & =\left(- \frac{b_{SU(2)}^{(0)}}{2\pi}\cdot
(l-l_{0})+\alpha_{2}^{-1}(l_{0})\right) \times\theta\left( l_{t}-l\right)
\\
& +\left(- \frac{b_{SU(2)}^{(0)} + \delta b_{SU(2)}}{2\pi}\cdot(l-l_{t})-\frac
{b_{SU(2)}^{(0)}}{2\pi}\cdot(l_{t} - l_{0})+\alpha_{2}^{-1}(l_{0})\right)
\times\theta\left( l-l_{t}\right) \\
\alpha_{1}^{-1}(l) & =\left(- \frac{b_{U(1)}^{(0)}}{2\pi}\cdot
(l-l_{0})+\alpha_{1}^{-1}(l_{0})\right) \times\theta\left( l_{t}-l\right) \\
& +\left(- \frac{b_{U(1)}^{(0)} + \delta b_{U(1)}}{2\pi}\cdot(l-l_{t})-\frac{b_{U(1)}%
^{(0)}}{2\pi}\cdot(l_{t} - l_{0})+\alpha_{1}^{-1}(l_{0})\right) \times
\theta\left( l-l_{t}\right)
\end{align}
where $l>l_{0} > 0$ denotes the RG time of the evolution, $l_{0}$ is a reference
value for the entry of the superparticles, and $l_{t}$ is the RG time of
the threshold. Here, $b^{(0)}_{G}$ refers to the usual one-loop MSSM beta functions,
and $\delta b_{G}$ refers to the threshold correction induced from coupling to the probe
D3-brane.
In the previous section we have seen that the beta function contributions $\delta b_{G}$
are not $SU(5)_{GUT}$ universal. We find that this leads to a distortion of unification
of the same size as two loop MSSM effects, but in the \textit{opposite} direction.
The contribution from such probe sectors therefore can actually help with precision unification.
To study the overall size of these effects, we now consider some representative
values for the threshold scale $M_{thresh} \sim M_{\cancel{CFT}}$. In principle,
we should also include the effects of the precise mass scales for all
superpartners. Rather than entangling this effect with the contribution from
just the probe sector, we simply take $l_{0} \sim 500$ GeV, with the
gauge couplings at the first threshold scale to be:%
\begin{equation}
(\alpha_{3}^{-1}(l_{0}),\alpha_{2}^{-1}(l_{0}),\alpha_{1}%
^{-1}(l_{0}))\sim(10,30,58).
\end{equation}
We note that with these values, the one-loop running
leads to a unified value of $\alpha_{GUT} \sim 0.05$ at
a scale $M_{GUT} \sim 2 \times 10^{16}$ GeV.
We define the GUT scale to be the scale at which
$\alpha_{2}^{-1}$ and $\alpha_{1}^{-1}$ unify, and we refer to this unified value
as $\alpha_{GUT}^{-1}$. We denote the percent mismatch between $\alpha_{3}$ and $\alpha_{GUT}$ by:%
\begin{equation}
\delta\equiv\frac{\alpha_{3}^{-1}(M_{GUT})-\alpha_{GUT}^{-1}}{\alpha
_{GUT}^{-1}}.
\end{equation}
For a threshold scale $M_{thresh} \sim 10^{13}$ GeV, this yields the
following numerical values for the various monodromy scenarios:
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|}\hline
$M_{thresh}\sim10^{13}$ GeV & $\delta b_{SU(3)}$ & $\delta b_{SU(2)}$ &
$\delta b_{U(1)}$ & $M_{GUT}$ (GeV) & $\alpha_{GUT}^{-1}$ & $\alpha_{3}^{-1}$
& $\delta$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & \multicolumn{1}{|c|}{$4.13$} & $4.$ & $4.05$ & $2\times10^{16}$
& $20.2$ & $20.0$ & $-1\%$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & \multicolumn{1}{|c|}{$4.$} & $4.$ & $4.$ & $2\times10^{16}$ &
$20.1$ & $20.1$ & $0\%$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $4.00$ & $3.65$ & $3.79$ & $2\times10^{16}$ & $20.7$ & $20.1$ &
$-3\%$\\\hline
$S_{3}$ & \multicolumn{1}{|c|}{$2.92$} & $2.75$ & $2.81$ & $2\times10^{16}$ &
$21.7$ & $21.4$ & $-1\%$\\\hline
$Dih_{4}^{(1)}$ & \multicolumn{1}{|c|}{$2.38$} & $2.08$ & $2.20$ &
$2\times10^{16}$ & $22.5$ & $22.1$ & $-2\%$\\\hline
$Dih_{4}^{(2)}$ & $2.16$ & $2.14$ & $2.15$ & $2\times10^{16}$ & $22.4$ &
$22.3$ & $-0.1\%$\\\hline
\end{tabular}
\ \ \ \ \ \ \ \ \ \ \ \ \ .
\end{equation}
From this, we see that in the one case where the beta functions are the same,
we retain one-loop unification, while in those cases with a small shift, there
is an improved agreement with precision unification. Of course, the value of $M_{GUT}$ quoted here does not include
other TeV and GUT scale thresholds, so this should really be taken as an order of magnitude estimate.
It is also of interest to compute the size of GUT distorting effects when the
threshold scale is pushed down to $\sim500$ GeV, which is close to the
maximal value allowed by present bounds. This is an interesting possibility to
see the largest possible distortion of gauge coupling unification in the
presence of such a probe sector. In this case, we obtain:%
\begin{equation}%
\begin{tabular}
[c]{|c|c|c|c|c|c|c|c|}\hline
$M_{thresh}\sim500$ GeV & $\delta b_{SU(3)}$ & $\delta b_{SU(2)}$ & $\delta
b_{U(1)}$ & $M_{GUT}$ (GeV) & $\alpha_{GUT}^{-1}$ & $\alpha_{3}^{-1}$ &
$\delta$\\\hline
$%
\mathbb{Z}
_{2}^{(1)}$ & $4.13$ & $4.$ & $4.05$ & $2\times10^{16}$ & $5.2$ & $4.4$ &
$-16\%$\\\hline
$%
\mathbb{Z}
_{2}^{(2)}$ & $4.$ & $4.$ & $4.$ & $2\times10^{16}$ & $5.0$ & $5.0$ &
$0\%$\\\hline
$%
\mathbb{Z}
_{2}\times%
\mathbb{Z}
_{2}$ & $4.00$ & $3.65$ & $3.79$ & $1\times10^{16}$ & $7.3$ & $5.1$ &
$-30\%$\\\hline
$S_{3}$ & $2.92$ & $2.75$ & $2.81$ & $2\times10^{16}$ & $11.5$ & $10.4$ &
$-9\%$\\\hline
$Dih_{4}^{(1)}$ & $2.38$ & $2.08$ & $2.20$ & $1\times10^{16}$ & $14.9$ &
$13.1$ & $-12\%$\\\hline
$Dih_{4}^{(2)}$ & $2.16$ & $2.14$ & $2.15$ & $2\times10^{16}$ & $14.3$ &
$14.2$ & $-1\%$\\\hline
\end{tabular}
\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ .
\end{equation}
As in the higher threshold case, the value of $M_{GUT}$ should really only be viewed as an order of magnitude estimate
since here we are neglecting various TeV and GUT scale thresholds. In this case,
the effects of the probe D3-brane sometimes induce a significant
\textquotedblleft overshoot\textquotedblright\ in the value of $\alpha_{3}$
versus $\alpha_{GUT}$ in the direction opposite to that expected from the
two-loop MSSM\ effects. It is interesting to note that even when we push the
scale of CFT\ breaking as low as $500$ GeV, the $S_{3}$ and $Dih^{(1)}_{4}$ monodromy scenarios
induce only an order $10 \%$ distortion in gauge coupling unification, while for the $Dih^{(2)}_{4}$ scenario,
the amount of distortion is on the order of $1 \%$.
Of course, to perform a more complete analysis of precision unification, we
should work to two loop order in the MSSM\ gauge couplings, include the
corresponding threshold corrections, and include all GUT scale
threshold corrections as well. This would involve a more detailed analysis
beyond the scope of the present paper. Nevertheless, the qualitative effect,
as well as its overall size, is clear: The presence of the probe D3-brane
provides all the necessary ingredients to improve precision unification.
\section{Supersymmetry Breaking\label{sec:SUSY}}
In this section we briefly consider the possibility that the hidden sector can
serve as the origin of supersymmetry breaking. In particular we argue that it
is natural in this setup to obtain a deformation of gauge-mediated
supersymmetry breaking. As discussed earlier, the breaking of the CFT is
naturally implemented by moving the D3-brane off the $7_{SM}$-brane. This displacement corresponds
to giving a vev to the operator $Z$ of the probe theory. We will also argue
that this procedure provides a natural mechanism for breaking supersymmetry. Once
the D3-brane moves off the $7_{SM}$-brane, the $3-7_{SM}$ strings which are
charged under the SM gauge group will serve as messengers in a gauge
mediated scenario, with the additional novelty that
the effective ``number'' of messenger fields is generically irrational but
computable. In addition, as already noted, the geometric profile of matter
fields in the internal directions of the compactification singles out the
Higgs fields and third generation as the dominant sources of coupling between
the probe and Standard Model. This leads to additional supersymmetry breaking
effects for the third generation and Higgs sectors, which can induce
potentially significant contributions to the $\mu,B\mu$, $A$-terms, and soft masses.
The organization of the rest of this section is as follows. First, we explain
why it is natural to expect supersymmetry breaking to occur in the probe brane
setup. Next, we discuss the resulting contributions from gauge mediated
supersymmetry breaking, and after this, the expected deformations to the Higgs
fields and third generation fields. Finally, we discuss how various
considerations from combining the probe and Standard Model sectors suggest
natural ranges of energy scales for the messenger scale of the model.
\subsection{CFT and SUSY Breaking on the D3-Brane}
In this subsection we discuss a geometrically natural way to implement
both conformal symmetry breaking and supersymmetry breaking on a probe D3-brane. The
basic scenario is as follows. We imagine that in the vicinity of the
configuration of intersecting seven-branes which realizes the Standard Model,
there is nearby a D3-brane. The position of this D3-brane in the geometry is
stabilized by various fluxes \cite{Martucci, FGUTSNC, Baumann:2010sx}. The general form of the
flux-induced superpotential will be a power series in the modes $Z_{i}$:%
\begin{equation}
W_{flux}=F_{i}Z_{i}+m_{ab}Z_{a}Z_{b}+\lambda_{ijk}Z_{i}Z_{j}Z_{k}+O(Z^{4}).
\end{equation}
The precise form of $W_{flux}$ will not concern us here. This deformation
persists even in the limit of non-compact seven-branes, because it can be
stated as a deformation of the topological B-model, independent of K\"{a}hler
data \cite{FGUTSNC}. Even so, the actual size of the deformation will depend on such data
since $W_{flux}$ is more accurately thought of as a section of a bundle. Once this
extra data is fixed, we see that $W_{flux}$ becomes quite dilute as we
decompactify the seven-brane (see also \cite{Marchesano:2009rz}). It is also
technically natural to treat $W_{flux}$ as a small perturbation to the
system. This is because $W_{flux}$ breaks additional flavor symmetries
of the probe sector. In what follows, we shall therefore ignore the
effects of $W_{flux}$ in determining the IR behavior of our probe theory.
Let us note that the presence of $W_{flux}$ will induce a further deformation of the
conformal sector. In \cite{FCFT} it was found that terms linear in $Z_{1}$ and
$Z_{2}$ do not lead to an interacting CFT in the infrared. See Appendix B of
\cite{D3gen} for a discussion of the fixed points obtained from terms
quadratic in the $Z_{i}$.
Thinking of $W_{flux}$ as a small perturbation, the presence of terms in $Z$ will
tend to attract the D3-brane close to the Yukawa point. Indeed, the fact that
such terms are irrelevant deformations of the CFT means that there will be a
supersymmetric vacuum located at $Z=0$, the origin of the Coulomb branch. For
generic enough $W_{flux}$, however, we can expect additional vacua close to the
Yukawa point. Indeed, when generic mass deformations for the $\mathcal{N}=2$
theory are switched on, for any point along the Coulomb branch of moduli
space there exists a suitable choice of $W_{flux}$ such that a metastable
vacuum at that prescribed value can be found \cite{Ooguri:2007iu, Auzzi:2010kv}. In our context with
$\mathcal{N}=1$ vacua, it is natural to expect that a similar situation occurs.
When $Z$ develops a non-zero vev, the D3-brane moves out on the Coulomb
branch, and conformal symmetry is also broken. Let us denote by
$M_{\cancel{CFT}}$ the energy scale at which conformal symmetry is broken.
Heuristically, we identify $\left\langle Z\right\rangle ^{1/\Delta}\sim
M_{\cancel{CFT}}$. We can therefore promote the effects of the vev to a
spurion chiral superfield $X$, which has scalar vev $M_{\cancel{CFT}}$.
Phrased in this way, we can also include the effects of supersymmetry breaking
in terms of the vev:%
\begin{equation}
\left\langle X\right\rangle =M_{\cancel{CFT}}+\theta^{2}F.
\end{equation}
Because $M_{\cancel{CFT}}$ specifies the scale of conformal symmetry breaking,
it follows that below this energy scale, the dynamics of the probe theory is
described by a theory with particle-like excitations. Note, however, that
since $\tau\sim O(1)$, this will be a strongly coupled theory. This
characteristic energy scale specifies the masses of some of the particle-like
excitations. For example, the mediator $3-7_{SM}$ strings, namely states
charged under the Standard Model gauge group will have mass $M_{\cancel{CFT}}%
$:%
\begin{equation}
M_{mess}\sim M_{\cancel{CFT}}.
\end{equation}
An interesting feature of probes with seven-brane monodromy is that some of
the GUT singlets of the probe sector will have masses far below
$M_{\cancel{CFT}}$, given instead by \cite{D3gen}:
\begin{equation}
M_{hid}\sim M_{\cancel{CFT}}\left( \frac{M_{\cancel{CFT}}}{M_{GUT}}\right)
^{\alpha}%
\end{equation}
for $\alpha\sim O(1)$ set by the details of seven-brane
monodromy.\footnote{For example, in weakly coupled models such as the $D_{4}$
probe theory $\alpha=1,2,3$, with the value set by the details of seven-brane
monodromy. Similar considerations are expected to hold in general
\cite{D3gen}.} Of course, the exact spectrum will also depend on various
D-terms and possible supersymmetry breaking effects.
\subsection{Gauge Mediated Contributions}
Introducing a source of supersymmetry breaking through the $Z$ field produces a
non-supersymmetric mass spectrum for the $3-7_{SM}$ strings. The fact that these
states are charged under the Standard Model gauge group means that the effects
of supersymmetry breaking in the probe sector will be communicated via gauge
interactions to the visible sector. In other words, there is a natural mode of SUSY breaking transmission
via gauge mediation.
To illustrate the basic ideas, we imagine that $Z$ develops a supersymmetry
breaking vev:%
\begin{equation}
\left\langle Z\right\rangle =\left( M_{mess}+\theta^{2}F\right)
^{\Delta_{IR}(Z)}.
\end{equation}
We now show that in the limit $F/M_{mess}^{2}\ll1$, it is possible to extract
the soft supersymmetry breaking contributions from gauge mediated
supersymmetry breaking. Our discussion closely follows \cite{Giudice:1997ni}.
To this end, recall that we have computed the shift in the SM gauge coupling
beta functions expected from coupling to the probe sector. Indeed, for%
\begin{equation}
b_{G}=-3\,\mathrm{Tr}(R_{IR}J_{G}J_{G}),
\end{equation}
with $G=SU(3),SU(2),U(1)$, we have defined the contribution from the messenger
sector to be given by a threshold contribution:%
\begin{equation}
\delta b_{G}=b_{G}-b_{G}^{(0)}.
\end{equation}
Let us now suppose that the $3-7_{SM}$ strings pick up a supersymmetric mass
$M_{mess}$. This is the messenger scale for our model. There is a Lagrangian
F-term contribution of the form
\begin{equation}
\delta\mathcal{L}=\frac{1}{8\pi}\operatorname{Im}\int d^{2}\theta\,\tau
_{G}\,\mathrm{Tr}_{G}\mathcal{W}^{\alpha}\mathcal{W}_{\alpha}-\frac{\delta
b_{G}}{16\pi^{2}}\operatorname{Re}\int d^{2}\theta\log M_{mess}\,\mathrm{Tr}%
_{G}\mathcal{W}^{\alpha}\mathcal{W}_{\alpha}. \label{running}%
\end{equation}
If we now imagine that the $3-7_{SM}$ mass spectrum is non-supersymmetric
$M_{mess}\rightarrow M_{mess}+\theta^{2}F$, we obtain a gaugino mass at the messenger scale:
\begin{equation}
m_{\lambda}=\delta b_{G}\left( \frac{\alpha_{G}}{4\pi}\right) \left(
\frac{F}{M_{mess}}\right) .
\end{equation}
Thus the effective number of messengers $N_{G}$ is, as expected, simply
\begin{equation}
N_{G}=\delta b_{G}.
\end{equation}
The interesting point in our case is that this number will generically be
irrational because of the CFT dynamics.
For the scalars there is the usual
\textquotedblleft two-loop" contribution from gauge mediation, where one of
the loops stands in for the CFT dynamics. As with the gauginos, it
is simple to read off the gauge mediated
contribution from the running of the gauge coupling. At the messenger scale,
this soft mass term is given by:
\begin{equation}
m_{\tilde{f}}^{2}=2\sum_{G}C^{G}_{2}(R_{\tilde{f}})N_{G}\left( \frac{\alpha_{G}%
}{4\pi}\right) ^{2}\left( \frac{F}{M_{mess}}\right) ^{2}%
\end{equation}
As before, this is what we expect as the contribution from $N_{G}$ messengers
for each gauge group $G$.
\subsection{Deformation of Gauge Mediation}
As noted before, there are additional couplings between the probe and the
visible sector. Geometrically, the main effects are from those Standard Model
fields which have maximal wavefunction profile at the point where the D3-brane
sits. In the flavor physics scenario of \cite{HVCKM, BHSV, FGUTSNC}, this
means there is a dominant coupling to the third generation and the Higgs fields.
Indeed, at the GUT scale, the Standard Model and probe become coupled via the
F-term deformation:
\begin{equation}
\delta W_{SM\oplus D3}=H_{u}\mathcal{O}_{H_{u}}+H_{d}\mathcal{O}_{H_{d}%
}+\overline{5}_{M}\mathcal{O}_{\overline{5}_{M}}+10_{M}\mathcal{O}_{10_{M}.}
\label{mixmix}%
\end{equation}
We can view the $\mathcal{O}$'s, which are operators in the CFT, as roughly
being made of composites of messenger fields charged under the MSSM gauge
group, whose beta function contributions we have already discussed. Integrating
these modes out below the $M_{\cancel{CFT}}$ scale will induce higher
dimension operators involving MSSM singlet operators from the D3-brane
sector coupled to the MSSM fields. In particular we generate an F-term
\begin{equation}
W_{eff}\supset\int d^{2}\theta\text{ }\mathcal{O}_{PQ}H_{u}H_{d}%
\end{equation}
where $\mathcal{O}_{PQ}$ is an MSSM neutral operator of the D3 brane theory
charged under the Peccei-Quinn symmetry. Here, we implicitly view this as a
higher dimension operator with suppression scale set by the mass scale of the
heavy messenger states, which is $M_{\cancel{CFT}}$. This term can provide a
mechanism for generating $\mu$ and $B\mu$ terms if $\mathcal{O}_{PQ}$ picks up
a vev. One can imagine that the $3-7_{SM}$ strings also communicate supersymmetry breaking
to the hidden sector via gauge mediated effects involving the $U(1)_{D3}$ of the probe theory.
In particular, various radiative corrections could naturally induce the analogue of
\textquotedblleft electroweak symmetry breaking\textquotedblright\ in the
hidden sector. Viewing $\mathcal{O}_{PQ}$ as a composite of $7_{flav}-3$ and
$3-7_{flav}$ modes, the basic idea is that these $3-7_{flav}$ strings can then
condense with supersymmetry breaking vevs which also break the PQ symmetry,
thus inducing $\mu$ and $B\mu$ terms. We are currently studying the dynamics
associated with such a scenario \cite{HVW}.
D-term corrections will also be generated. Some such corrections are already
present from gauge mediated contributions. Indeed, in minimal gauge mediation,
the soft scalar masses can be computed by tracking the dependence of
wave-function renormalization factors on a supersymmetry breaking spurion
\cite{Giudice:1997ni, ArkaniHamed:1998kj}. The presence of the F-term couplings
$\Psi_{R} \cdot \mathcal{O}_{\overline{R}}$ introduce additional messenger/matter couplings,
involving mainly the third generation and Higgs fields. These introduce
additional contributions to the soft masses and $A$-terms. Let us note that
because the dominant coupling is to the third generation, we do not expect
such contributions to induce large flavor changing neutral currents in the
first two generations.
As one can see from this discussion, there is a rich set of possibilities
for potential soft supersymmetry breaking patterns motivated by D3-brane considerations.
\subsection{Comments on the CFT Breaking/Messenger Scale}
Up to this point, we have treated the CFT breaking scale (which is the
messenger scale) as an undetermined parameter. There are two considerations
which allow us to narrow down the values of this parameter. First, we observe
that if we go all the way to the infrared fixed point, then the Standard Model
Yukawa couplings involving the ``third generation'' are all irrelevant. In this case, it would be
more appropriate to view the mode identified at the GUT scale as a ``second generation'' up type quark as
an effectively ``third generation'' field with a tuned top quark Yukawa. We do not entertain this possibility here.
Keeping the CFT breaking scale as low as possible is appealing in the gauge mediated framework, because the CFT
breaking and supersymmetry breaking scales can then be roughly the same.
To see how low we can push the CFT breaking scale, let us consider the amount of suppression expected for the top quark Yukawa coupling
in running from the GUT scale down to the CFT breaking scale. To illustrate the main idea, suppose
that the dimension of the $5_{H} \times 10_{M} \times 10_{M}$ term is
of order $3 + \epsilon \sim 3.1$, which is numerically similar to the values obtained in many of the scenarios.
In this case, demanding that the Yukawa coupling is sufficiently large imposes
the constraint:
\begin{equation}
\lambda_{5 \times 10 \times 10} \sim\left( \frac{M_{\cancel{CFT}}}{M_{GUT}}\right)^{0.1} \gtrsim 10^{-1}
\end{equation}
where here, we allow an order $10$ fine-tuning in the GUT scale value of the Yukawa.
This leads to a lower bound on the CFT breaking scale, of order:
\begin{equation}
M_{\cancel{CFT}}\gtrsim10^{3}\,\,\text{TeV.}%
\end{equation}
While this is close to the TeV scale, it is clearly above it. The precise lower bound depends on
the amount of tuning of the Yukawas at the GUT scale, and the particular probe scenario in question.
In other words, considerations from the visible sector, in particular that flavor physics
be generated correctly at the GUT scale, disfavors TeV scale values for the
hidden sector.
Confining our attention to the most conservative possibility where flavor is generated
at the GUT scale, retaining sufficiently large Yukawas then suggests increasing
the CFT breaking scale to an intermediate value, of order $10^{9} - 10^{13}$ GeV. This
is also in the range which is most helpful for precision unification in most
of the scenarios we have considered. Additionally, this is the range of values
which are most natural for D3-brane induced baryon and dark matter genesis
scenarios \cite{D3gen}. In this regime, there is also a
natural hierarchy between the mass scale of the messenger particles, and the
mass scale of the GUT singlets \cite{D3gen}. This fits nicely with the fact that
if the $\mu$-term is generated at lower energy scales, it should involve the
dynamics of light degrees of freedom below the CFT breaking scale.
Let us note that especially in the single D3-brane $Dih^{(2)}_{4}$ probe
theory, the amount of suppression to the Yukawas is quite mild. In this case, the dimension of the $5_{H} \times 10_{M} \times 10_{M}$
operator is $\sim 3 + 0.04$, while that of the $\overline{5}_{H} \times \overline{5}_{M} \times 10_{M}$ has dimension of order
$\sim 3 + 0.26$. Since the bottom quark is significantly lighter than the top quark,
one could also contemplate a moderate to low $\tan \beta$ scenario in which the bottom quark
is instead identified with the GUT scale ``second generation'' mode. Indeed, in the
flavor scenario of \cite{HVCKM}, the ratio of the second and third generation masses at the GUT scale is only
of order $\alpha_{GUT} \sim 0.05$. This might allow significantly lower CFT breaking scales
for the $Dih^{(2)}_{4}$ scenario, which in principle could be accessible at the LHC.
Note that the amount of distortion to unification in the $Dih^{(2)}_{4}$ scenario is also smaller than
in the other cases, and begins to counteract two-loop MSSM effects when the CFT breaking scale is quite low.
Of course, this would also require us to revisit various assumptions about flavor physics. Though a full study
of such a scenario is beyond the scope of the present work, this is perhaps the most phenomenologically exciting possibility.
\section{Conclusions \label{sec:CONC}}
E-type Yukawa points are required in order to generate a top quark mass in
F-theory GUTs. D3-brane probes of such E-points are then a very well-motivated
extension of the Standard Model. In this paper we
have studied the effects of the Standard Model on such a probe sector, and
conversely, the effects of the probe on the Standard Model. We have presented
evidence for the existence of a strongly interacting conformal fixed point for
this system, and moreover, have shown that various properties of this system,
such as the infrared R-symmetry, the scaling dimensions of operators, and the
effects of the probe on the running of the gauge coupling constants are all
computable. This analysis reveals some remarkable features, notably:
\begin{itemize}
\item Such fixed points exist.
\item The MSSM fields develop
small anomalous dimensions, and thus maintain their identity in the IR.
\item The presence of the probe sector states coming in complete GUT
multiplets can help with precision unification.
\end{itemize}
We have also seen that the
D3-brane can serve as a natural source of supersymmetry breaking in
which the gauginos and first two generations have a spectrum similar to gauge
mediated supersymmetry breaking, with additional deformations of the
third generation and Higgs sectors. In the remainder of this section we discuss
various future directions.
In this paper we have mainly focused on the regime of energy scales
between the CFT breaking scale $M_{\cancel{CFT}}$ and the GUT scale. Below
this scale, various fields will develop vevs, and they will affect the low
energy phenomenology. Some studies of metastable vacua for theories with
underlying $\mathcal{N} = 2$ supersymmetry have been considered in the
literature. It would clearly be of interest to extend this analysis to the
types of systems studied here. Along these lines, it is also natural to
consider how supersymmetry breaking on the D3-brane would influence the
visible sector. We have seen that the probe D3-brane system leads to a pattern
of supersymmetry breaking terms which mainly deforms the third generation and
Higgs fields away from what is expected in minimal gauge mediation.
Determining the full range of possibilities, and the associated phenomenology
would clearly be of interest.
We have also seen that in spite of appearances, qualitative as well
as quantitative features of the probe/MSSM couplings can be extracted from this
scenario. In particular, we have studied the running of the gauge couplings,
and the consequences for gauge coupling unification. After moving onto the
Coulomb branch, the probe D3-brane contributes another $U(1)_{D3}$ gauge group
factor, which is strongly coupled at the CFT breaking scale. In particular,
the value of the holomorphic gauge coupling $\tau_{D3}(M_{\cancel{CFT}})
\sim\exp(2 \pi i /3)$ near an $E_{6}$ or $E_{8}$ point. At lower scales, this
$U(1)_{D3}$ is expected to be broken. It would be interesting to see whether an
exact computation of the running down to the $U(1)_{D3}$ breaking scale could also
be performed. It would also be of interest to determine the exact type of
kinetic mixing expected between the abelian factor of the Standard Model and
this $U(1)_{D3}$ gauge group factor.
Another feature of the probe sector is that it contains operators with the
same gauge quantum numbers as the usual MSSM Higgs fields. Most of the usual
headaches with electroweak symmetry breaking and its naturalness stem from
the vector-like nature of the Higgs fields. Though somewhat removed from the
considerations of the present paper, an intriguing possibility is that the
operators of the probe sector could effectively function as Higgs fields. This
can be interpreted as dissolving the D3-brane into a finite size instanton in
the visible sector seven-brane.
\newpage
\section*{Acknowledgements}
We thank N. Arkani-Hamed, C. C\'{o}rdova, D. Green, Z. Komargodski, P. Langacker, D. Poland, and S.-J. Rey for helpful
discussions. JJH and BW thank the Harvard high energy theory group for
generous hospitality during part of this work. CV thanks the CTP at MIT for
hospitality during his sabbatical leave. The work of JJH is supported by NSF
grant PHY-0969448. The work of CV is supported by NSF grant PHY-0244821. The
work of BW is supported by the US Department of Energy under grant DE-FG02-95ER40899.
|
\section{Introduction}
Throughout this paper we fix a smooth, connected, projective curve $C$ of genus at least $2$.
For a complex Lie group $G$ we denote by
$\mathcal{M}_C(G)$ the moduli stack of principal $G$-bundles and by $\mathcal{L}$ the ample line bundle that generates the Picard group $\mathrm{Pic}(\mathcal{M}_C(G))$.
The spaces $H^0(\mathcal{M}_C(G),\mathcal{L}^{l})$ of generalized $G$-theta functions of level $l$ are well-known for classical Lie groups but
less understood for exceptional Lie groups.
Let $G_2$ be the smallest exceptional Lie group: the group of automorphisms of the complex Cayley algebra.
Our aim is to relate the space of generalized $G_2$-theta functions $H^0(\mathcal{M}_C(G_2),\mathcal{L})$ of level one
to other spaces of generalized theta functions associated to classical Lie groups.
Using the Verlinde formula, which gives the dimension of the space of generalized $G$-theta functions for any simple and simply-connected Lie group $G$,
we observe a numerical coincidence: $$\dim H^0(\mathcal{M}_C(SL_2),\mathcal{L}^3)=2^g \dim H^0(\mathcal{M}_C(G_2),\mathcal{L}).$$ In addition, we link together $\dim H^0(\mathcal{M}_C(G_2),\mathcal{L})$ and $\dim H^0(\mathcal{M}_C(SL_3),\mathcal{L})$.
Our aim is to give a geometric interpretation of these dimension equalities.
According to the Borel-De Siebenthal classification \cite{BorelDeSiebenthal},
the groups $SL_3$ and $SO_4$ appear as the two maximal subgroups of $G_2$ among the connected subgroups of $G_2$ of maximal rank.
We define two linear maps by pull-back of the corresponding extension maps: on the one hand
$$H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SL_3),\mathcal{L})$$
and on the other hand using the isogeny $SL_2\times SL_2\rightarrow SO_4$:
$$H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow
H^0(\mathcal{M}_C(SL_2), \mathcal{L})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{3}).$$
These maps take values in the invariant part by the duality involution for the first map and by the action of
$2$-torsion elements of the Jacobian for the second one.
We denote these invariant spaces by
$H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$ and
$\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 $ respectively.
Using the natural isomorphism proved in \cite{BNR}:
$$H^0(\mathcal{M}_C(SL_3),\mathcal{L})^*\simeq H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)$$
where $\Theta =\{L \in \mathrm{Pic}^{g-1}(C)\ | \ h^0(C,L)>0 \}$, we prove the following theorem:
\begin{theoremalph}
The linear map
$$ \Phi :H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$$
obtained by pull-back of the extension map $\mathcal{M}_C(SL_3)\rightarrow \mathcal{M}_C(G_2)$ is surjective for a general curve and is an isomorphism when the genus of the curve equals $2$.
\end{theoremalph}
A curve is said satisfying the cubic normality when the multiplication map $\mathrm{Sym}^3 H^0(\mathcal{M}_C(SL_2), \mathcal{L}) \rightarrow H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{3})$ is surjective.
Using an explicit basis of $H^0(\mathcal{M}_C(SL_2), \mathcal{L}^2)$ described in \cite{BeauvilleII}, we prove the theorem:
\begin{theoremalph}
The linear map
$$\Psi : H^0(\mathcal{M}_C(G_2),\mathcal{L}_{G_2})\rightarrow
\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 $$
obtained by pull-back of the extension map $\mathcal{M}_C(SO_4)\rightarrow \mathcal{M}_C(G_2)$
is an isomorphism for a general curve satisfying the cubic normality.
\end{theoremalph}
\begin{nota}
We use the following notations:
\begin{itemize}
\item
$C$ a smooth, connected, projective curve of genus $g$ at least 2,
\item
$G$ a connected and simply-connected simple complex Lie group,
\item
$\mathcal{M}_C(G)$ the moduli stack of principal $G$-bundles on $C$,
\item
$K_C$ the canonical bundle on $C$ ,
\item
$\mathrm{Pic}^{g-1}(C)$ the Picard group parametrizing line bundles on $C$ of degree $g-1$,
\item
$h^0(X,\mathcal{L})=\dim H^0(X,\mathcal{L})$.
\end{itemize}
\end{nota}
\section{Principal $G_2$-bundles arising from vector bundles of rank two and three}
\subsection{The octonions algebra}
Let $\OO$ be the complex algebra of the octonions, $\mathrm{Im}(\OO)$ the $7$-dimensional subalgebra of the imaginary part of $\OO$ and $\mathcal{B}_0=(e_1, \dots, e_7)$ the canonical basis of $\mathrm{Im}(\OO)$(see \cite{Baez}).
The exceptional Lie group $G_2$ is the group of automorphisms of the octonions.
In Appendix \ref{appendix The octonion algebra} we give the multiplication table in $\mathrm{Im}(\OO)$ by the Fano diagram and we introduce another basis $\mathcal{B}_1=(y_1, \dots, y_7)$ of $\mathrm{Im}(\OO)$
so that $\langle y_1, y_2, y_3\rangle$ and $\langle y_4, y_5, y_6 \rangle$ are isotropic and orthogonal and $y_7$ is orthogonal to both of these subspaces. This basis is defined in Appendix \ref{appendix The octonion algebra}, as well as two other basis obtained by permutation of elements of $\mathcal{B}_1$:
$\mathcal{B}_2=(y_2,y_3,y_4,y_5,y_6,y_1,y_7)$ and $\mathcal{B}_3=(y_1,y_2,y_4,y_5,y_3,y_6,y_7)$.
In the following paragraphs, we use local sections of rank-$7$ vector bundles satisfying multiplication rules of $\mathcal{B}_2$ or $\mathcal{B}_3$.
\subsection{Principal $G_2$-bundles admitting a reduction}
We introduce the notion of non-degenerated trilinear form on $\mathrm{Im}(\OO)$ as Engel did it in \cite{Engel1} and\cite{Engel2}.
A trilinear form $\omega$ on $\mathrm{Im}(\OO)$ is said non-degenerated if the associated bilinear symmetric form $B_\omega$ is non-degenerated where $B_\omega(x,y) =\omega(x,\cdot,\cdot)\wedge \omega(y,\cdot,\cdot)\wedge \omega(\cdot,\cdot,\cdot)$ $\forall x, y \in \mathrm{Im}(\OO)$.
\begin{lem}
Giving a principal $G_2$-bundle is equivalent to giving a rank-$7$ vector bundle with a non-degenerated trilinear form.
\end{lem}
\begin{proof}
Let $P$ be a principal $G_2$-bundle on $C$ and $V$ the associated rank-$7$ vector bundle and let $\omega$ be any non-degenerated trilinear form on $\mathrm{Im}(\OO)$.
By construction, $V$ has a reduction to $G_2$, \textit{i.e.} it exists a section $\sigma : C\rightarrow GL_7/G_2$.
The Lie group $G_2$ is the stabilizer $\mathrm{Stab}_{SL_7}(\omega)$ under the action of $SL_7$. Besides, under the action of $GL_7$, $\mathrm{Stab}_{GL_7}(\omega_0)\simeq G_2 \times \mathbb{Z}/3\mathbb{Z}$. Then
$$ \sigma : C\stackrel{\sigma}{\rightarrow} GL_7/G_2 \twoheadrightarrow GL_7/(G_2 \times \mathbb{Z}/3\mathbb{Z})\simeq GL_7/ \mathrm{Stab}_{GL_7}(\omega) \simeq \mathrm{Orb}_{GL_7}(\omega).$$
In addition, the orbit $\mathrm{Orb}_{GL_7}(\omega)$ is the set of all the non-degenerated trilinear form on $\mathrm{Im}(\OO)$.
So, $V$ is fitted with a non-degenerated trilinear form.
Reciprocally any rank-$7$ vector bundle fitted with a non-degenerated trilinear form defines a $G_2$-vector bundle.
\end{proof}
For a principal $G_2$-bundle, we use the non-degenerate trilinear form $\omega$ on $\mathcal{V}$, locally defined by
$\omega(x,y,z)=-\mathrm{Re}[(xy)z]$.
According to the Borel-De Siebenthal classification (see \cite{BorelDeSiebenthal}), $SL_3$ and $SO_4$ are, up to conjugation, the two maximal subgroups of $G_2$ among the connected subgroups of $G_2$ of maximal rank.
Using the inclusion $G_2 \subset SO_7=SO(\mathrm{Im}(\OO))$
both of the following lemma describe the rank-$7$ vector bundle (and the non-degenerate trilinear form) associated to a principal $G_2$-bundle which admits either a $SL_3$-reduction or a $SO_4$-reduction.
Note that $\mathrm{M}_C(SO_4)$ has two connected components distinguished by the second Stiefel Whitney class.
We only make here explicit computations with regards to the connected component $\mathrm{M}_C^+(SO_4)$ of $\mathrm{M}_C(SO_4)$ containing the trivial bundle.
\subsubsection{Principal $G_2$-bundles arising from rank-$3$ vector bundles}
\begin{lem}
\label{Lem Inclusion SL3 dans G2 and associated vector bundle}
Let $E$ be a rank-$3$ vector bundle with trivial determinant and let
$E(G_2)$ be his associated principal $G_2$-bundle and $\mathcal{V}$ be his associated rank-$7$ vector bundle.
Then, $\mathcal{V}$ has the following decomposition and the local sections basis $\mathcal{B}_2$ is adapted to this decomposition:
$$\mathcal{V}=E\oplus E^* \oplus \mathcal{O}_C.$$
The non-degenerate trilinear form $\omega$ is defined by the following local conditions:
\begin{enumerate}
\item
$\Lambda^3 E\simeq \Lambda^3 E^* \simeq \C$ and
$\omega (y_2,y_3,y_4)=\omega (y_5,y_6,y_1) =-\sqrt{2}$.
\item
On $E\times E^*\times\mathcal{O}_C$:
$$\omega (y_2,y_5,y_7)=\omega (y_3,y_6,y_7)=\omega (y_4,y_1,y_7) =i,$$
\item
All other computation, not obtainable by permutation of the previous triplets, equals zero.
\end{enumerate}
\end{lem}
\begin{proof}
Under the action of $SL_3$, $\mathrm{Im}(\OO)$ decomposes into $SL_3$-modules:
$$
\begin{array}{rcl}
\mathrm{Im}(\OO)& =& \langle y_2 , y_3 , y_4 \rangle \oplus \langle y_5 , y_6 , y_1\rangle \oplus \langle y_7\rangle\\
&=&\C^3 \oplus (\C^3)^* \oplus \C
\end{array}
$$
where $\{y_2 , y_3 , y_4, y_5 , y_6 , y_1, y_7\}$ is the basis $\mathcal{B}_2$ defined in Appendix \ref{appendix The octonion algebra};
$\langle y_2 , y_3 , y_4\rangle$ is an isotropic subspace of dimension $3$, dual of $\langle y_5 , y_6 , y_1\rangle$.
So, the $7$-rank vector bundle $\mathcal{V}$ associated to a rank-$3$ vector bundle $E$ is:
$$
\begin{array}{rl}
\mathcal{V} = & E \stackrel{SL_3}{\times}(\C^3 \oplus (\C^3)^* \oplus \C),\\
\mathcal{V}= & E\oplus E^* \oplus \mathcal{O}_C.
\end{array}
$$
Evaluations given for $\omega$ are deduced from Table \ref{tabular B2} of Appendix \ref{appendix The octonion algebra}.
\end{proof}
\subsubsection{Principal $G_2$-bundles arising from two rank-$2$ vector bundles}
The following lemma makes explicit the vector bundle associated to a principal $G_2$-bundle extension of an element of $\mathrm{M}_C^+(SO_4)$, using the surjective map $\mathrm{M}_C(SL_2)\times \mathrm{M}_C(SL_2)\twoheadrightarrow \mathrm{M}_C^+(SO_4)$.
\begin{lem}
\label{Lem Inclusion SL2*SL2 dans G2 and associated vector bundle}
Let $E, F$ be two rank-$2$ vector bundles of trivial determinant.
Denote by $(E,F)$ the associated principal $SO_4$-bundle, $P$ the associated principal $G_2$-bundle and $\mathcal{V}$ the associated rank-$7$ vector bundle.
Then, $\mathcal{V}$ has the following decomposition and the local sections basis $\mathcal{B}_3$ is adapted to this decomposition:
$$\mathcal{V}=E^*\otimes F \oplus \mathrm{End}_0(F).$$
The non-degenerate trilinear form $\omega$ on $\mathcal{V}$
is defined by the following local conditions:
\begin{enumerate}
\item
On $(E^*\otimes F)^3$ and on $(E^*\otimes F )\times(\mathrm{End}_0(F))^2$, $\omega$ is identically zero,
\item
On $(E^*\otimes F)^2 \times\mathrm{End}_0(F)$:
$$
\begin{array}{rcl}
\omega (y_2,y_4,y_3)&=\omega (y_5,y_1,y_6) &=\sqrt{2},\\
\omega (y_4,y_1,y_7) &=\omega (y_2,y_5,y_7)&=i.
\end{array}
$$
\item
$\Lambda^3\mathrm{End}_0(F) \simeq \C$ and $\omega (y_3,y_6,y_7) =i,$
\item
All other computation, not obtainable by permutation of the previous triplets, equals zero.
\end{enumerate}
\end{lem}
\begin{proof}
The groups $SO_4$ and $SL_2\times SL_2$ are isogenous.
Under the action of $SO_4$, $\mathrm{Im}(\OO)$ has the following decomposition:
$$
\begin{array}{rcl}
\mathrm{Im}(\OO)&=& \langle y_1,y_2,y_4,y_5\rangle \oplus \langle y_3,y_6,y_7\rangle, \\
&\simeq& M^*\otimes N \stackrel{\bot}{\oplus} \mathrm{End}_0(N)
\end{array}
$$
where $M,N$ are $2$-dimensional; an element $\overline{(A,B)}$ of $SO_4$ ($A,B \in SL_2$) acts on $ M^*\otimes N$ by $A\otimes B$ and by conjugation $\mathrm{End}_0(N)$.
So, the rank-$7$ vector bundle $\mathcal{V}$ associated to $(E,F)(G_2)$, when $E,F$ are two rank-$7$ vector bundle of trivial determinant, is:
$$\mathcal{V}=E^*\otimes F \oplus \mathrm{End}_0(F).$$
Evaluations given for $\omega$ are deduced from Table \ref{tabular B3} of Appendix \ref{appendix The octonion algebra}.
\end{proof}
\section{Equalities between dimensions of spaces of generalized theta functions}
Here are some dimension counts, using the Verlinde Formula, to calculate
$h^0(\mathcal{M}_C(G_2),\mathcal{L})$ and $h^0(\mathcal{M}_C(SL_2),\mathcal{L}^3)$.
\begin{prop}
\label{Prop h0(G2,L) h0(SL2,L3)}
Dimensional equalities between the following spaces of generalized theta functions
occur:
\begin{align}
h^0(\mathcal{M}_C(G_2),\mathcal{L}) &= \left(\frac{5+\sqrt{5}}{2}\right)^{g-1} +\left(\frac{5-\sqrt{5}}{2}\right)^{g-1},
\label{h0(G2,L)}\\
h^0(\mathcal{M}_C(SL_2),\mathcal{L}^3)&= 2^g\left[\left(\frac{5+\sqrt{5}}{2}\right)^{g-1}+
\left(\frac{5-\sqrt{5}}{2}\right)^{g-1}\right] ,\label{h0(SL2,L3)}\\
\text{so, }\quad h^0(\mathcal{M}_C(SL_2),\mathcal{L}^3)&=2^g h^0(\mathcal{M}_C(G_2),\mathcal{L}). \label{h0(SL2,L3)=2g h0(G2,L)}
\end{align}
\end{prop}
\begin{proof}
See Appendix \ref{appendix h0(G2,L)} and \ref{appendix h0(SL2,L3)}
\end{proof}
\section{Surjectivities and isomorphisms between $H^0(\mathcal{M}_C(G_2),\mathcal{L})$ and $H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$}
To avoid confusion, we sometimes specify the group $G$ in the notation of the generator of the Picard group $\mathcal{L}$ writing $\mathcal{L}_G$.
Let $ H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$ be the invariant part of $H^0(\mathcal{M}_C(SL_3),\mathcal{L})$ by the duality involution:
the eigenspace of $H^0(\mathcal{M}_C(SL_3),\mathcal{L})$
associated to the eigenvalue $1$ under the natural involution $E \mapsto \sigma (E)=E^*$.
We consider the extension map
$\mathrm{i} : \mathcal{M}_C(SL_3)\rightarrow \mathcal{M}_C(G_2)$
which associates to a rank-$3$ vector bundle of trivial determinant the associated principal $G_2$-bundle.
The pull-back $\mathrm{i}^*(\mathcal{L}_{G_2})$ equals $\mathcal{L}_{SL_3}$.
\begin{thm}
The extension map
$\mathrm{i} : \mathcal{M}_C(SL_3)\rightarrow \mathcal{M}_C(G_2)$
induces by
pull-back a linear map between the following spaces of generalized theta functions:
$$H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SL_3),\mathrm{i}^*(\mathcal{L}_{G_2})).$$
This map takes values in $ H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$.
We denote by $\Phi$ this map:
$$ \Phi :H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+.$$
\end{thm}
\begin{proof}
\begin{enumerate}
\item
The pull-back $\mathrm{i}^*(\mathcal{L}_{G_2})$ equals $\mathcal{L}_{SL_3}$.Indeed, let $E$ be a rank-$3$ vector bundle.
By Lemma \ref{Lem Inclusion SL3 dans G2 and associated vector bundle}, the rank-$7$ vector bundle associated to $E$ is $E\oplus E^* \oplus \mathcal{O}_C$.
We study the commutative diagram:
\begin{diagram}
\mathcal{M}_C(G_2) & \rTo^{\rho_1} & \mathcal{M}_C(SL_7)\\
\uTo^{\mathrm{i}} & & \uTo_{\rho_2}\\
\mathcal{M}_C(SL_3) & \rTo^{\rho_3} & \mathcal{M}_C(SL_3)\times \mathcal{M}_C(SL_3)
\end{diagram}
where $\mathrm{i}$ and $\rho_1$ are maps of extension of group of structure and $\forall E \in \mathcal{M}_C(SL_3) ,$
$\rho_3(E)=(E, E^*)$ and $\forall (E,F) \in \mathcal{M}_C(SL_3)\times \mathcal{M}_C(SL_3)$,
$\rho_2 (E,F) =E \oplus F \oplus \mathcal{O}_C$.
By Proposition $2.6$ of \cite{LaszloSorger}, applied with $SL_7$, $G_2$
and the irreducible representation $\rho_1$ of highest weight $\varpi_1$, the Dynkin index of $d(\rho_1)$ equals $2$.
Therefore, $\rho_1^*(\mathcal{D}_{SL_7})=\mathcal{L}_{G_2}^{2}$ where $\mathcal{D}_{SL_7}$ the determinant bundle generator of $\mathrm{Pic}(\mathcal{M}_C(SL_7))$ (see \cite{KNR} and \cite{LaszloSorger}).
In addition, by the same proposition,
$\rho_2^*(\mathcal{D}_{SL_7})=\mathcal{L}_{SL_3} \boxtimes \mathcal{L}_{SL_3}$
and
$\rho_3^*(\rho_2^*(\mathcal{D}_{SL_7}))=\mathcal{L}_{SL_3}^{ 2}$.
So, $\mathrm{i}^*(\mathcal{L}_{G_2})^{2}=\mathcal{L}_{SL_3}^{ 2}$ so that $\mathrm{i}^*(\mathcal{L}_{G_2})=\mathcal{L}_{SL_3}$ since the Picard group
$\mathrm{Pic}(\mathcal{M}_C(SL_3))$ is isomorphic to $\mathbb{Z}$.
\item
We show that the image of the linear map $\Phi$ is contained in $H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$.
The morphism $\mathrm{i}$ is $\sigma$-invariant: for all $ E\in \mathcal{M}_C(SL_3)$, the $G_2$-principal bundles $E(G_2)$ and $E^*(G_2)$ are isomorphic.
Indeed, the Weyl group $ W(SL_3)$ is contained in $W(G_2)$ since so are there normalizer; and $ W(SL_3)$ is a subgroup of $W(G_2)$ of index $2$.
We consider $\overline{g}$ in the Weyl group $W(G_2)\backslash W(SL_3)$ and $g \in G_2$ a representative of the equivalence class of $\overline{g}$. Then, $g\notin SL_3$.
Let $C_g$ be the inner automorphism of $G_2$ induced by $g$. As the subalgebra $\got{sl}_3$ of $\got{g}_2$ corresponds to the long roots and as each element of the Weyl group $W(G_2)$ respects the Killing form on $\got{g}_2$, $C_g(SL_3)$ is contained in $SL_3$. The restriction of $C_g$ to $SL_3$ is then an exterior automorphism of $SL_3$
which we call
$\alpha : SL_3 \rightarrow SL_3$.
This automorphism exchanges the two fundamental representations of $SL_3$.
So, $\alpha$ induces an automorphism $\widetilde{\alpha}$ on $\mathcal{M}_C(SL_3)$ such that, $\forall E\in \mathcal{M}_C(SL_3)$, $\widetilde{\alpha}(E)=E^*$. Consider the following commutative diagram, where $\widetilde{C}_{g}$ is
the inner automorphism given by $g$:
\begin{diagram}
\mathcal{M}_C(SL_3) &\rInto &\mathcal{M}_C(G_2) \\
\dTo^{\widetilde{\alpha}}&&\dTo_{\widetilde{C_{g}}}\\
\mathcal{M}_C(SL_3) & \rInto & \mathcal{M}_C(G_2).
\end{diagram}
Then, $\forall E\in \mathcal{M}_C(SL_3)$,
$$E^*(G_2)=\widetilde{\alpha}(E)(G_2)= \widetilde{C_{g}}(E(G_2))\simeq E(G_2)$$
since $ \widetilde{C_{g}}$ is an inner automorphism.
Thus, $\mathrm{i}(E)$ and $\mathrm{i}(\sigma(E))$ are isomorphic.
The $\sigma$-invariance of $\mathrm{i}$ implies that the image of $\Phi$ is contained in one of the two eigenspaces of $H^0(\mathcal{M}_C(SL_3),\mathcal{L})$.
As $\sigma^*(\mathcal{L}_{SL_3})\simeq\mathcal{L}_{SL_3}$, which is the isomorphism which implies identity over the trivial bundle, we get $\sigma(\mathrm{i}^*(\mathcal{L}_{G_2}))=\sigma^*(\mathcal{L}_{SL_3})=\mathcal{L}_{SL_3}=\mathrm{i}^*(\mathcal{L}_{G_2})$.
Thus, the image of $\Phi$ is contained in $H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$, the eigenspace relative to the eigenvalue $1$.
\end{enumerate}
\end{proof}
We remind two points of vocabulary:
an \textit{even theta-characteristic} $\kappa$ on a curve $C$ is an element $\kappa$ of $\mathrm{Pic}^{g-1}(C)$ such that $\kappa \otimes \kappa = K_C$ and $h^0(C, \kappa)$ is even ;
a curve $C$ is said \textit{without effective theta-constant} if $h^0(C, \kappa)=0$ for all even theta-characteristic $\kappa$.
The set of all even theta-characteristics is named
$\Theta^{\mathrm{even}}(C)$.
\begin{thm}
\label{Thm Application Phi}
The linear map $\Phi :H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$
\begin{enumerate}
\item
\label{thm item Phi surjective for C without eff theta-constant}
is surjective when the curve $C$ is without effective theta-constant.
\item
is an isomorphism if the genus of $C$ equals $2$.
\end{enumerate}
\end{thm}
\begin{proof}
\begin{enumerate}
\item
Let $C$ be a curve without effective theta-constant and consider the following diagram:
\begin{diagram}
\mathcal{M}_C(G_2) & \rTo^{\rho_1} & \mathcal{M}_C(SL_7)\\
\uTo^{\mathrm{i}} & \ruTo_{\rho} & \\
\mathcal{M}_C(SL_3) & &
\end{diagram}
We introduce the element $\Delta_\kappa$ defined for each even theta-characteristic~$\kappa$:
$$ \Delta_\kappa =\{P\in \mathcal{M}_C(SL_7)\ | \ h^0(C, P(\C^7)\otimes \kappa) > 0\}. $$
These $ \Delta_\kappa$ are Cartier divisors, so they define, up to a scalar, an element of $ H^0(\mathcal{M}_C(SL_7), \mathcal{L})$.
The image $\Phi(\rho_1^*(\Delta_\kappa))=\rho^*(\Delta_\kappa)$ is
$$
\begin{array}{rl}
\rho^*(\Delta_\kappa)=&\{E\in \mathcal{M}_C(SL_3)\ |\ h^0(C,E\oplus E^* \oplus \mathcal{O}_C) \otimes \kappa) >0 \},\\
=&\{E\in \mathcal{M}_C(SL_3)\ |\ h^0(C,E\otimes \kappa)+h^0(C,E^*\otimes \kappa)+ h^0(C,\kappa) >0 \},\\
=&\{E\in \mathcal{M}_C(SL_3)\ |\ h^0(C,E\otimes \kappa)+h^0(C, E^*\otimes \kappa) >0 \},\\
&\text{because $C$ is without effective theta-constant,}\\
=&\{E\in \mathcal{M}_C(SL_3)\ |\ 2h^0(C,E\otimes \kappa) >0 \}, \text{ by Serre duality}.\\
\end{array}
$$
Thus, $\rho^*(\Delta_\kappa)=2H_\kappa$
where $H_\kappa:=\{ E \in \mathcal{M}_C(SL_3) \ | \ h^0(C,\mathcal{E}\otimes \kappa)>0\}$.
Therefore, to show the surjectivity of $\Phi$, it suffices to show that $\{H_\kappa\ | \ \kappa \in \Theta^{\mathrm{even}}(C)\}$
generates $H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$.
We consider
$$\Theta =\{L \in \mathrm{Pic}^{g-1}(C)\ | \ h^0(C,L)>0 \}$$ and the natural map between the spaces $H^0(\mathcal{M}_C(SL_3),\mathcal{L})^*$ and $H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)$.
By Theorem $3$ of \cite{BNR}, this map is an isomorphism and, besides, it is
equivariant for the two involutions on $H^0(\mathcal{M}_C(SL_3),\mathcal{L})^*$ and $H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)$ (respectively $E \mapsto E^*$ and $L\mapsto K_C\otimes L^{-1}$).
So, the components $(+)$ and $(-)$ of each part are in correspondence: $ H^0(\mathcal{M}_C(SL_3),\mathcal{L})^*_+$ is isomorphic to $H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)_+$.
Denote by $\varphi$ this isomorphism
$\varphi : \PP H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+ \stackrel{\sim}{\rightarrow} \PP H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)_+^* $. For all even theta-characteristic $\kappa$, the image $\varphi(H_\kappa)$ is $\varphi_{3\Theta}(\kappa)$ where $\varphi_{3\Theta}$ is the following map:
$$\varphi_{3\Theta} : \mathrm{Pic}^{g-1}(C) \rightarrow \PP H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)_+^*=|3 \Theta|^*_+ .$$
The set $\{ \varphi_{3\Theta}(\kappa)\ | \ \kappa \in \Theta^{\mathrm{even}}(C)\}$
generates $|3 \Theta|^*_+$.
Indeed, in the following commutative diagram
\begin{diagram}
& & |4\Theta|^*_+ \\
&\ruDashto^{\varphi_{4\Theta}} & \dDashto \\
\mathrm{Pic}^{g-1}(C) & \rDashto^{\varphi_{3\Theta}} & | 3\Theta|^*_+ ,
\end{diagram}
the map $|4\Theta|^*_+ \dashrightarrow |
3\Theta|^*_+$ is surjective because it is induced by the inclusion
$D \in H^0(C,3\Theta)_+ \mapsto D+\Theta \in H^0(C,4\Theta)_+$. In addition, by
\cite{SermanKopeliovichPauly}, when $C$ is without effective theta-constant,
$\{\varphi_{4\Theta}(\kappa)\ |\ \kappa \in \Theta^{\mathrm{even}}(C)\}$
is a base of $|4\Theta|^*_+$ (the number of even theta-characteristics equals $2^{g-1}(2^g+1)$
which equals the linear dimension of $|4\Theta|^*_+$).
Thus,
$\{ \varphi_{3\Theta}(\kappa)\ |\ \kappa\in \Theta^{\mathrm{even}}(C)\}$
generates $|3 \Theta|^*_+$ and $\{H_\kappa\ |\ \kappa\in \Theta^{\mathrm{even}}(C)\}$ generates the space $H^0(\mathcal{M}_C(SL_3),\mathcal{L})_+$.
As we have shown that $H_\kappa$ equals $\Phi(\rho_1(\Delta_\kappa))$ for all $\kappa$ even theta-characteristic, the map $\Phi$ is surjective.
\item
By \cite{BNR}, the dimension of $ H^0(\mathcal{M}_C(SL_3),\mathcal{L})^*_+$ equals the dimension of $H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)_+$, that is $\frac{3^g+1}{2}$.
When the genus of $C$ is $2$, the dimension of $H^0(\mathrm{Pic}^{g-1}(C), 3 \Theta)_+$ equals $5$ which is also the dimension of $H^0(\mathcal{M}_C(G_2),\mathcal{L})$ by Proposition \ref{Prop h0(G2,L) h0(SL2,L3)}.\eqref{h0(G2,L)}.
A curve of genus $2$ is without effective theta-constant. So, by this dimension equality and by the point \eqref{thm item Phi surjective for C without eff theta-constant} of the theorem, $\Phi$ is an isomorphism when the genus of $C$ equals $2$.
\end{enumerate}
\end{proof}
\section{Isomorphisms between spaces of generalized $G_2$-theta functions and generalized $SL_2$-theta functions}
Let $\mathrm{JC}[2]$ be the group of $2$-torsion elements of the Jacobian: $\mathrm{JC}[2]=\{\alpha \in \mathrm{Pic}^0(C)\ |\ \alpha \otimes \alpha =\mathcal{O}_C\}$.
This group acts on $\mathcal{M}_{SL_2} \times \mathcal{M}_{SL_2}$ : for $ \alpha\in \mathrm{JC}[2]$ and
$(E,F)\in \mathcal{M}_{SL_2} \times \mathcal{M}_{SL_2} $, we associate $(E\otimes \alpha,F\otimes \alpha)$.
Let $\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 $ be the invariant part
of $\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]$
under the action of the element of the Jacobian group $\mathrm{JC}[2]$.
We study the subgroup $SO_4$ of $G_2$, which is isogenous to $SL_2\times SL_2$, and the linear map induced by pull-back by the extension map $\mathrm{j} : \mathcal{M}_C(SO_4)\rightarrow \mathcal{M}_C(G_2)$
which associates to two rank-$2$ vector bundle of trivial determinant the associated principal $G_2$-bundle.
The pull-back $j^*(\mathcal{L}_{G_2})$ equals $\mathcal{L}_{SL_2}\boxtimes\mathcal{L}_{SL_2}^{3}$
\begin{thm}
\label{Theorem definition Psi}
The extension map
$\mathrm{j} : \mathcal{M}_C(SO_4)\rightarrow \mathcal{M}_C(G_2)$
induces by pull-back a linear map between the following spaces of generalized theta functions:
$$ H^0(\mathcal{M}_C(G_2),\mathcal{L})\rightarrow H^0(\mathcal{M}_C(SO_4),j^*(\mathcal{L}_{G_2})).$$
This map takes values in $\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 $.
We denote by $\Psi$ this map:
$$\Psi : H^0(\mathcal{M}_C(G_2),\mathcal{L}_{G_2})\rightarrow
\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 $$
\end{thm}
\begin{proof}
\begin{enumerate}
\item
Consider the following commutative diagram:
\begin{diagram}
\mathcal{M}_C(G_2) & \rTo^{\rho_1} & \mathcal{M}_C(SL_7)\\
\uTo^{\mathrm{j}} & & \uTo_{\rho_4} \\
\mathcal{M}_C(SL_2)\times \mathcal{M}_C(SL_2)& \rTo^{(f_1,f_2)} &\mathcal{M}_C(SL_4)\times \mathcal{M}_C(SL_3)\\
\end{diagram}
where $\mathrm{j}$ and $\rho_1$ are the extension maps, $f_1(M,N)=M^*\otimes N $,
$f_2(M,N)= \mathrm{End}_0(N)$ and $\rho_4(A,B)=A\oplus B$.
As in the previous section, we calculate explicitly $j^*(\mathcal{L}_{G_2})$.
Let $\mathcal{D}_{SL_7}$ be the determinant bundle of $\mathcal{M}_C(SL_7)$ and $\mathrm{pr}_1$ and $\mathrm{pr}_2$
the canonical projections of
$SL_2\times SL_2$. We get
$$f_1^*(\mathcal{L}_{SL_4})=\mathrm{pr}_1^*(\mathcal{L}_{SL_2})^{2}\otimes \mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{2}=\mathcal{L}_{SL_2}^{2} \boxtimes \mathcal{L}_{SL_2}^{2}$$
and according to Table $B$ of \cite{SorgerModuliofPrincipalbundles}:
$$f_2^*(\mathcal{L}_{SL_3})=\mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{4},$$
since $f_2$ is associated to the adjoint representation of $SL_2$, which has Dynkin index $4$.
$$
\begin{array}{rrcl}
&\rho_4^*(\mathcal{D})&=&\mathcal{L}_{SL_4}\boxtimes \mathcal{L}_{SL_3},\\
\text{so}\quad&j^*(\mathcal{L}_{G_2}^{2})&=&(f_1,f_2)^*(\mathcal{L}_{SL_4}\boxtimes \mathcal{L}_{SL_3}),\\
&&=&f_1^*(\mathcal{L}_{SL_4})\otimes f_2^*(\mathcal{L}_{SL_3}),\\
&&=&[\mathrm{pr}_1^*(\mathcal{L}_{SL_2})^{2}\otimes \mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{2}]\otimes
\mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{4},\\
&j^*(\mathcal{L}_{G_2}^{2})&=&\mathrm{pr}_1^*(\mathcal{L}_{SL_2})^{2}\otimes \mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{6},\\
\text{so}\quad&j^*(\mathcal{L}_{G_2})&=&\mathrm{pr}_1^*(\mathcal{L}_{SL_2})\otimes \mathrm{pr}_2^*(\mathcal{L}_{SL_2})^{3}.
\end{array}
$$
We get
$$j^*(\mathcal{L}_{G_2})=\mathcal{L}_{SL_2}\boxtimes\mathcal{L}_{SL_2}^{3}.$$
\item
The morphism $\mathrm{j} : \mathcal{M}_{SL_2} \times \mathcal{M}_{SL_2}\rightarrow \mathcal{M}_{G_2}$ is invariant under the action of $JC[2]$:
$(E,F)\in \mathcal{M}_{SL_2} \times \mathcal{M}_{SL_2}$ and $\alpha \in JC[2]$ then the rank-$7$ vector bundle associated by $\mathrm{j}$ to $(E\otimes \alpha,F\otimes \alpha)$ is
$$
\begin{array}{rl}
&(E\otimes \alpha)^* \otimes (F\otimes \alpha) \oplus \mathrm{End}_0(F\otimes \alpha)\\
=&E^*\otimes F\otimes \alpha^*\otimes \alpha \oplus \mathrm{End}_0(F\otimes \alpha),\\
=&E\otimes F\oplus \mathrm{End}_0(F)= \mathrm{j}(E,F).
\end{array}
$$
Therefore, the image of $\Psi$ is contained in the expected vector space
$[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})]_0$.
\end{enumerate}
\end{proof}
Before going further in the study of the morphism $\mathrm{j}$, we compare the dimensions of the involved sets.
\begin{lem}
\label{Lem dim[H0(MSL2,LSL2) and H0(MSL2, LSL2 3)]0 = dim H0(MG2,LG2)}
The dimension of the space $\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0$ equals the dimension of $H^0(\mathcal{M}_C(G_2),\mathcal{L}_{G_2})$.
\end{lem}
\begin{proof}
By Proposition \ref{Prop h0(G2,L) h0(SL2,L3)},\eqref{h0(SL2,L3)=2g h0(G2,L)}, we notice the remarkable following relation:
\begin{align}
\label{Equation h Msl2, L3 = 2^g h Mg2, L}
2^g h^0\left(\mathcal{M}_C(G_2),\mathcal{L}\right)=h^0(\mathcal{M}_C(SL_2),\mathcal{L}^{3}).
\end{align}
So, as $h^0(\mathcal{M}_C(SL_2),\mathcal{L})=2^g$ (see \cite{BeauvilleI}), we get
$$
\begin{array}{ll}
\dim \left(\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0 \right)&\\
=\frac{1}{2^{2g}}\times h^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\times h^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3}) & \text{because}\ |\,JC[2]\,|=2^{2g},\\
=\frac{1}{2^{2g}}\times 2^{g}\times 2^{g}h^0(\mathcal{M}_C(G_2), \mathcal{L}_{G_2}) &\text{by}\ \eqref{Equation h Msl2, L3 = 2^g h Mg2, L},\\
=h^0(\mathcal{M}_C(G_2), \mathcal{L}_{G_2}).&
\end{array}$$
\end{proof}
All the following results are based on the \textit{cubic normality conjecture}. Its statement is:
\begin{conj}
\label{Conj surjection HO(MSL2,L) in H0(MSL2, L3)}
For a general curve $C$, the multiplication map
$$\eta : \mathrm{Sym}^3 H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}) \rightarrow H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})$$
is surjective.
\end{conj}
When the previous map $\eta$ is surjective, we say that the curve $C$ satisfies \textit{cubic normality}.
\begin{prop}
\label{Prop normalité cubique vraie}
Cubic normality holds for
all curve of genus $2$, all non hyper-elliptic curve of genus $3$ and all curve of genus $4$ without effective theta-constant.
\end{prop}
\begin{proof}
For a curve of genus $2$, $\mathcal{M}_C(SL_2)$ is isomorphic to $\PP^3$
and $\mathcal{L}_{SL_2}$ to $\mathcal{O}(1)$ (see \cite{NarasimhanRamanan}).
A non-hyper-elliptic curve of genus $3$ is a Coble quartic (see \cite{NR}).
The cubic normality is true in both of these cases.
For a general curve of genus $4$ without effective theta-constant,
cubic normality is proved in Theorem
$4.1$ of \cite{OxburyPauly}.
\end{proof}
When this conjecture is true, we get this theorem:
\begin{thm}
\label{Thm : Conj implique isom}
Let $C$ be a curve of genus at least $2$ without effective theta-constant and satisfying the cubic normality and let $\Psi$ be the map defined in Theorem \ref{Theorem definition Psi}:
$$\Psi : H^0(\mathcal{M}_C(G_2),\mathcal{L}_{G_2})\rightarrow
\left[H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2})\otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}_{SL_2}^{3})\right]_0. $$
\begin{enumerate}
\item
\label{item Proposition Conj normalité cubique implique Psi isomorphisme}
The map $\Psi$ is an isomorphism,
\item
\label{fait 2 H0(MG2, LG2) engendré par les rho1 (delta kappa)}
The space of generalized $G_2$-theta functions $H^0(\mathcal{M}_C(G_2), \mathcal{L})$ is linearly generated by the divisors
$\rho_0^*(\Delta_\kappa)$ for $\kappa$ even theta-characteristic,
where $\rho_0$ is the extension morphism
$$\rho_0 : \mathcal{M}_C(G_2) \rightarrow \mathcal{M}_C(SO_7).$$
\end{enumerate}
\end{thm}
\begin{proof}
\begin{enumerate}
\item
According to the dimension equality proved in Lemma \ref{Lem dim[H0(MSL2,LSL2) and H0(MSL2, LSL2 3)]0 = dim H0(MG2,LG2)},
it suffices to prove the surjectivity of $\Psi$.
Denote by $V$ the vector space $H^0(\mathcal{M}_C(SL_2), \mathcal{L})$.
Using the notations of \cite{BeauvilleII}, we associate to each even theta-characteristic $\kappa$ an element
$d_\kappa$ of $H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{\otimes 2})$ and an element $\xi_\kappa $ of $V\otimes V$.
For each even theta-characteristic $\kappa$, $d_\kappa$ is the section of $H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{\otimes 2})$ such that $D_\kappa$ is the divisor of the zeros of $d_\kappa$, where $D_\kappa=\{S \in M_C(SL_2)\ |\ h^0(C ,\mathrm{End}_0(S)\otimes \kappa)>0\}$.
Consider the following maps:
$$
\begin{array}{rcl}
\rho_0^* :&H^0( \mathcal{M}_C^+(SO_7), \mathcal{L})&\longrightarrow H^0(\mathcal{M}_C(G_2),\mathcal{L})\\
\text{and}\quad \beta :&[ V \otimes V \otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{2})]_0&\longrightarrow [V \otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{3}]_0
\end{array}
$$
where $\beta(A,B,D)=(A,BD)$.
For any even theta-characteristic $\kappa$, the image $\Psi(\rho_0^*(\Delta_\kappa))$ equals $\beta(\xi_\kappa\otimes d_\kappa)$. Indeed, $\Psi$ is induced by:
$$
\begin{array}{rrcl}
\mathrm{j}:&\mathcal{M}_C(SL_2)\times\mathcal{M}_C(SL_2) &\rightarrow &\mathcal{M}_C(G_2) \\
&(E,F) & \mapsto &\mathrm{Hom}(E,F) \oplus \mathrm{End}_0(F);
\end{array}
$$
the pull-back $\Psi(\rho^*_0(\Delta_\kappa))$ is the sum of two divisors:
$$
\begin{array}{rcl}
\Delta_1 & =&\{(E,F)\in \mathcal{M}_C(SL_2)\times\mathcal{M}_C(SL_2) \ |\ h^0(C,\mathrm{End}_0(F) \otimes \kappa) >0 \},\\
\Delta_2 & =&\{(E,F)\in \mathcal{M}_C(SL_2)\times\mathcal{M}_C(SL_2) \ |\ h^0(C, \mathrm{Hom}(E,F) \otimes \kappa >0 \}.
\end{array}
$$
In addition, $\mathcal{O}(\Delta_1)=\mathcal{O}_C \boxtimes \mathcal{L}^{2}$ and $\mathcal{O}(\Delta_2)= \mathcal{L} \boxtimes \mathcal{L}$ (see \cite{BeauvilleII}) and more precisely:
$$\Delta_1 = \text{Zeros}(d_\kappa) \text{ et } \Delta_2= \text{Zeros}(\xi_\kappa) .$$
When the curve $C$ is of genus at least $2$ without effective theta-constant,
it is proved in \cite{BNR} that the map
$$
\begin{array}{rrcl}
\varphi_0^* : & \mathrm{Sym}^2 V &\longrightarrow& H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{2})\\
& \xi_\kappa &\mapsto& d_\kappa
\end{array}
$$
is an isomorphism. We identify $\mathrm{Sym}^2 V$ with the invariant space of $V\otimes V$ under the involution $a\otimes b\mapsto b\otimes a$.
By Theorem $1.2$ and Proposition A.$5$ of \cite{BeauvilleII}, the set $\{d_\kappa\ |\ \kappa \in\Theta^{\mathrm{even}}(C) \}$ is a basis of $H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{2})$ and $\{\xi_\kappa\ |\ \kappa \in\Theta^{\mathrm{even}}(C)\} $ is a basis of $\mathrm{Sym}^2 V$.
Then, the vector space $[ V \otimes V \otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{2})]_0$ is generated by $\{\xi_\kappa \otimes d_\kappa \ |\ \kappa \in\Theta^{\mathrm{even}}(C) \}$.
Thus, to prove the surjectivity of the map $\Psi$, it is sufficient to show the surjectivity of the map $\beta$.
Consider the following diagram:
\begin{diagram}
V\otimes H^0(\mathcal{M}_C(SL_2),\mathcal{L}^{2})&\rTo &H^0(\mathcal{M}_C(SL_2),\mathcal{L}^{3})\\
\uTo &&\uTo_{\eta} \\
V \otimes V\otimes V &\rTo & \mathrm{Sym}^3V.
\end{diagram}
With the hypothesis of cubic normality, the map $\eta$ is surjective. Therefore the map $V\otimes H^0(\mathcal{M}_C(SL_2),\mathcal{L}^{2}) \rightarrow H^0(\mathcal{M}_C(SL_2),\mathcal{L}^{3})$ is also surjective. By restriction to invariant sections under the action of $JC[2]$, $\beta$ is surjective.
The map $\Psi$ is thus an isomorphism.
\item
The point \eqref{fait 2 H0(MG2, LG2) engendré par les rho1 (delta kappa)} is a consequence of the previous facts:
for each element $\kappa \in \Theta^{\mathrm{even}}(C)$, the image of $\rho_0^*(\Delta_\kappa)$ by $\Psi$ is $\xi_\kappa \otimes d_\kappa$.
As $\{\xi_\kappa \otimes d_\kappa \ |\ \kappa \in \Theta^{\mathrm{even}}(C)\}$ generates $[ V \otimes V \otimes H^0(\mathcal{M}_C(SL_2), \mathcal{L}^{2})]$, the set $\{\rho_0^*(\Delta_\kappa)\ |\ \kappa \in \Theta^{\mathrm{even}}(C)\}$ generates $H^0(\mathcal{M}_C(G_2), \mathcal{L})$.
\end{enumerate}
\end{proof}
\begin{rem}
By Proposition \ref{Prop normalité cubique vraie}
the linear map $\Psi$ is an isomorphism for
each curve of genus $2$, each non hyperelliptic curve of genus $3$ and each curve of genus $4$ without effective theta-constant.
\end{rem}
|
\section{Introduction}
It is as well theoretically proven as experimentally tested that the exploitation of entangled states improves our ability to manage information in several ways \cite{hororev}. The development of quantum information theory is definitely due to this fundamental result \cite{nielsen}. However, it has been recently discovered that even multipartite separable states can play a relevant role in performing better-than-classical communication and information protocols \cite{dattaqc,piani,singapore}. In general, then, the usefulness of a particular state of a quantum system for such tasks can be traced back to the presence of internal correlations among parts of the system, i.e., to the knowledge that an observer Alice, probing a subsystem $A$, gains about the state of another subsystem $B$, controlled by another observer Bob, and vice versa.
The quantitative and qualitative evaluation of such correlations and the proper discrimination of their nature --- classical versus quantum --- stand as open problems. Several quantifiers of non-classicality of correlations have been introduced in literature \cite{zurek,HV,terhal,mid,deficit,moelmer,modi,dakic,gap,rossignoli} (not to be confused with the non-classicality of quantum optical states \cite{nonclasscorr}), but there are still neither clear criteria of faithfulness or {\it bona-fide}-ness for them, nor a well established hierarchy of reliability. In spite of that, they are heavily used in current research \cite{usediscord,usemid,ferraro,discordGauss,operdiscord,mista,provapovm,luo,alber,cinesi_sensati,nuovodiscord,james,galve,luofu}. The most popular one is by far the \emph{quantum discord} \cite{zurek,HV}, a measure of non--classical correlations which goes beyond entanglement and whose definition has an immediate interpretation in information theory: discord equals the difference of the total correlations between two subsystems $A$ and $B$ before and after a local measurement process is performed on one of them. The quantum discord admits at least three other operational interpretations, in contexts ranging from thermodynamics to communication protocols, such as the quantum state merging \cite{operdiscord}. The evaluation of the quantum discord is a hard task from a computational point of view, implying an optimization of the conditional entropy between Alice and Bob, ${\cal S}(A|B)$, over all local (generalized) measurements on one party, which is often obtainable by numerical methods only. A closed analytical solution is known in the case of arbitrary two--mode Gaussian states \cite{discordGauss}, under the restriction of Gaussian local measurements. Narrowing our overview to two--qubit states, an analytical expression of discord has been derived in particular for the subclass of so-called $X$--states \cite{luo,alber,cinesi_sensati,nuovodiscord}, and a successful attempt to generalize this procedure has not been advanced yet (to the best of our knowledge), apart from an upper bound quantity for the discord defined in \cite{cinesi_sensati}.
The difficulty in calculating quantum discord motivated the introduction of alternative measures of non-classical correlations.
In particular, the \emph{geometric discord} \cite{dakic} is one such a measure, which quantifies the amount of non-classical correlations of a state in terms of its minimal distance from the set of genuinely classical states. The geometric discord involves a simpler optimization and is easily computable analytically for general two--qubit states. However, its relationship with the original quantum discord is not entirely clear at the present stage.
In this work, we present an algorithm to calculate quantum discord for general two--qubit states. First, we obtain an explicit and simplified expression for the conditional entropy, exploiting the Bloch representation of the density matrix; then, we employ new variables that allow us to set the optimization conditions in a closed form. Finally, we associate them to constraints over the eigenvalues of the statistical ensemble obtained after the measurement process. Our approach qualifies as the most efficient and reliable way to evaluate quantum discord for arbitrary states of two qubits \cite{noterespect}.
Exploiting our algorithm, we perform a detailed numerical exploration of the Hilbert space of two--qubit states to compare quantum discord and the geometric discord as quantifiers of non-classical correlations. We shed light on the relationship between these two quantities by identifying the states that extremize geometric discord at fixed quantum discord (and vice versa). We are motivated by the aim of establishing a reliable hierarchy of non-classical states based on physically and mathematically consistent criteria. We find that, interestingly, the quantum discord of a two--qubit state can never exceed its (normalized) geometric discord. In analogy with the study of maximal entanglement \cite{nemoto}, we find that the notion of maximal non-classical correlations is measure--dependent: Therefore, the feasibility in a specific experimental realization will determine which, among the various classes of maximally quantumly correlated states, are the most suitable ones to be employed for applications.
The paper is organized as follows. Section \ref{sec1} provides an introduction to quantum discord and geometric discord (hereafter denoted by ${\cal D}$ and $D_G$, respectively), presenting definitions, main properties and a summary of the results obtained in the literature. In Section \ref{sec2}, we pursue an analytical approach to the calculation of quantum discord for general two--qubit states, eventually recasting the optimization problem into a system of two transcendental equations, whose solution specifies the local measurement that minimizes the conditional entropy. Section \ref{sec3} concerns the comparison between quantum discord and geometric discord by using the results of the previous sections; we identify the classes of states with maximal and minimal geometric discord at fixed quantum discord. Finally, Section \ref{sec4} recalls the main results of our work and suggests further issues worthy of investigation.
\section{Basic definitions}\label{sec1}
\subsection{Quantum discord}\label{secQD}
One of the lessons we can apprehend from Quantum Mechanics is that the measurement process disturbs the state in which a physical system is set. This differs from what happens in the classical scenario, hence it is possible to conclude that the disturbance induced by a measurement on a state is a good evidence of its ``quantumness''. Now, let us suppose to have a bipartite system in a certain state and make a measurement on one of its subsystems. We can analyze the nature of the internal correlations of the system in such a state by studying in which way they are affected by the measurement. Information Theory provides the tools to accomplish this task.
As a starting point, let us consider the case of information stored in two classical probability distributions. The quantity expressing the total amount of correlations between two random variables $X$ and $Y$, assuming values $\{x_i\},\{y_j\}$ with probability $\{p_i\},\{q_j\}$, is the \emph{mutual information} ${\cal I}$, defined as
\begin{eqnarray}\label{ixy}
{\cal I}(X:Y)={\cal H}(X)+{\cal H}(Y)-{\cal H}(X,Y),
\end{eqnarray}
where ${\cal H}(X)=-\sum _i p_i \log_2 p_i$ is the Shannon entropy associated to the random variable $X$; consequently ${\cal H}(Y)=-\sum _j q_j \log_2 q_j$ is the entropy for $Y$ while ${\cal H}(X,Y)$ is the joint entropy of $X$ and $Y$.
Following Bayesian rules, we can retrieve equivalent formulations, ${\cal I}(X:Y)={\cal I}(Y:X)={\cal J}(X:Y)= {\cal J}(Y:X)$, where
\begin{eqnarray}
{\cal J}(X:Y)&=&{\cal H}(X)-{\cal H}(X|Y)\,,\\
{\cal J}(Y:X)&=&{\cal H}(Y)-{\cal H}(Y|X)\,.
\end{eqnarray}
Here the conditional entropy is straightforwardly defined as ${\cal H}(X|Y)={\cal H}(X,Y)-{\cal H}(Y)$, and represents how much uncertainty (ignorance) we have on $X$ given the value of $Y$ (and vice versa for ${\cal H}(Y|X)$).
In the quantum scenario, we consider a bipartite system $AB$ described by a density matrix $\rho \equiv \rho_{AB}$, and subsystems $A,B$ with marginal density matrices $\rho_A, \rho_B$. The mutual information can be used once more to quantify the total correlations between $A$ and $B$. The quantum analogue of the expression (\ref{ixy}) straightforwardly reads
\begin{eqnarray}
{\cal I}(A:B)={\cal S}(A)+{\cal S}(B)-{\cal S}(A,B),
\end{eqnarray}
where ${\cal S}(A)=-\text{Tr}[\rho_A \log_2 \rho_A]$ is the von Neumann entropy of subsystem $A$, and equivalently ${\cal S}(B)=-\text{Tr}[\rho_B \log_2 \rho_B]$, ${\cal S}(A,B)=-\text{Tr}[\rho \log_2 \rho]$. On the other hand, if we try to define a quantum version of ${\cal J}$ we have
\begin{eqnarray}
{\cal J}(A:B)={\cal S}(A)-{\cal S}(A|B),
\end{eqnarray}
which is an ``ambiguous'' quantity \cite{zurek,HV}, since the conditional quantum entropy ${\cal S}(A|B)$ is clearly depending on which observable we have measured on $B$. Using the same notation of \cite{alber}, we recall that a von Neumann measurement (from now on just measurement) on $B$ projects the system into a statistical ensemble $\{p_k,\rho_k\}$, such that
\begin{eqnarray}\label{measure}
\rho\rightarrow \rho_k =\frac{(I_A\otimes P_{Bk})\rho (I_A\otimes P_{Bk} )}{p_k},
\end{eqnarray}
where
\begin{eqnarray}
p_k&=& \text{Tr}[\rho (I_A\otimes P_{Bk} )] \\\nonumber
P_{Bk}&=&V \Pi _k V^\dagger\\ \nonumber
\Pi _k &= &|k \rangle\langle k|,\ \ k=0,1\\\nonumber
V &\in& SU(2). \nonumber
\end{eqnarray}
We remind that in this case study the use of generalized positive-operator-valued-measurements (POVMs) is not required, since it has been proven in \cite{provapovm} that for two--qubit states the optimal measurement for the conditional entropy is always a projective one. We can say that the amount of truly classical correlations is expressed by the mutual information obtained adopting the least disturbing measurement \cite{HV}:
\begin{eqnarray}
{\cal C}(A:B)&=&\text{max}_{\{P_{Bk}\}}{\cal J}(A:B) \\
&=&{\cal S}(A)-\text{min}_{\{P_{Bk}\}}\sum_k p_k {\cal S}(A|B_{\{P_{Bk}\}}).\nonumber
\end{eqnarray}
Consequently, the amount of genuinely quantum correlations, called \emph{quantum discord}, is given by \cite{zurek}
\begin{eqnarray}\label{discord}
&&{\cal D}(A:B)={\cal I}(A:B)- {\cal C}(A:B)\\\nonumber
&&\quad = {\cal S}(B)- {\cal S}(A,B) + \text{min}_{\{P_{Bk}\}}\sum_k p_k {\cal S}(A|B_{\{P_{Bk}\}}).
\end{eqnarray}
We note that quantum discord is not symmetric:
\begin{eqnarray}
{\cal S}(A)-{\cal S}(A|B)&\neq& {\cal S}(B)-{\cal S}(B|A) ;
\end{eqnarray}
performing the measurement on Alice's subsystem rather than on Bob's one is perfectly legitimate, but it returns in general a different value of discord. See e.g. \cite{gap,mista} for a discussion about the implications of it. It is immediate to verify that not only entangled states, but almost all separable states have a non--vanishing quantum discord \cite{ferraro}, i.e. are affected by the measurement process, thus exhibiting some pretty quantum properties. In the case of pure bipartite states, the discord reduces to the marginal entropy of one of the two subsystems and therefore to the canonical measure of entanglement. Quantum discord for two--qubit states is normalized to one.
\subsection{Geometric discord}
Recently, it has been argued that the experienced difficulty of calculating quantum discord can be coped, for a general two--qubit state, with the introduction of its geometrized version, hereby just called \emph{geometric discord} \cite{dakic}.\\
Let us suppose, to be coherent with section \ref{secQD}, to have a bipartite system $AB$ and to make a measurement on $B$. As we have remarked, almost all (entangled or separable) states are disturbed by the measurement; however, there is a subclass of states which is invariant and presents zero discord. It is the class of the so--called \emph{classical--quantum} states \cite{piani}, whose elements have a density matrix of this form
\begin{eqnarray}\label{cq}
\rho = \sum _i p_i \rho _{Ai}\otimes |i\rangle \langle i | ,
\end{eqnarray}
where $p_i$ is a probability distribution, $\rho_{Ai}$ is the marginal density matrix of $A$ and $\{|i\rangle\}$ is an orthonormal vector set. A classical--quantum state is not affected by a measurement on $B$ in any case.
Letting $\Omega$ be the set of classical--quantum two--qubit states, and $\chi$ be a generic element of this set, the geometric discord $D_G$ is defined as the distance between the state $\rho$ and the closest classical--quantum state. In the original definition \cite{dakic}, the (squared) Hilbert--Schmidt distance is adopted. Recalling that $||A ||_2^2=\text{Tr}[A A^T]$ is the square of the Hilbert--Schmidt norm of a matrix $A$, the geometric discord has been introduced as
\begin{eqnarray}\label{dgdef}
D_G(\rho)=\text{min}_{\chi \in \Omega}||\rho -\chi ||_2^2.
\end{eqnarray}
It is possible to obtain an explicit closed expression of $D_G$ for two--qubit states. First, one needs to express the $4\times 4$ density matrix of a two--qubit state in the so--called Bloch basis \cite{verstraete}:
\begin{eqnarray}\label{bloch}
\rho&=& \frac14 \sum_{i,j=0}^3 R_{ij} \sigma_i \otimes \sigma_j \\
\nonumber &=& \frac 14\bigg(I_{4\times 4}+\sum_{i=1}^3 x_i\sigma_i \otimes I_{2\times 2} \\ \nonumber
& & \quad +\sum_{j=1}^3 y_j I_{2\times 2}\otimes \sigma_j+\sum_{i,j=1}^3 t_{ij} \sigma _i\otimes\sigma_j\bigg),
\end{eqnarray}
where $R_{ij}=\text{Tr}[\rho(\sigma_i\otimes \sigma_j)]$, $\sigma_0=I_{2\times 2}$, $\sigma _i$ ($i=1,2,3$) are the Pauli matrices, $\vec{x}=\{x_i\},\vec{y}=\{y_i\}$ are the three--dimensional Bloch vectors associated to the subsystems $A,B$, and $t_{ij}$ denote the elements of the correlation matrix $T$. Then, it is shown in \cite{dakic} that the geometric discord is given by
\begin{eqnarray}\label{dgformula}
D_G(\rho)= \frac 14(||\vec y \vec y^T||_2 + ||T||_2^2 -k),
\end{eqnarray}
with $k$ being the largest eigenvalue of the matrix $\vec y \vec y^T+ T^TT$ (in case of measurement on Alice, one needs to replace $\vec{y}$ with $\vec{x}$ and $T^TT$ with $TT^T$). An alternative formulation for the geometric discord has been provided in \cite{luofu}. It is easy to see that $D_G$ is not normalized to one: its maximum value is $1/2$ for two--qubit states, so it is natural to consider $2D_G$ as a proper measure for a comparison with the quantum discord ${\cal D}$.
\section{Quantum discord for two--qubit states}\label{sec2}
\subsection{General setting}
Even though quantum discord has an apparently simple definition \cite{zurek}, the practice reveals that its explicit evaluation is hard to accomplish. In this paper we restrict our attention to two--qubit states. An analytical algorithm has been proposed for the subclass of states with maximally mixed marginals (described by five real parameters) in \cite{luo}. Also, an extension to states spanned by seven real parameters, called $X$--states because of the peculiar form of their density matrix (with vanishing elements outside the leading diagonal and the antidiagonal), has been introduced in \cite{alber}, and amended by \cite{nuovodiscord}. Here, we attempt to generalize the procedure to the entire class of two--qubit states.
First, we consider that performing local unitary transformations we can recast the density matrix for an arbitrary two--qubit state, Eq.~(\ref{bloch}), in the Bloch normal form \cite{verstraete,luo}
\begin{eqnarray}\label{simpler}
\rho&=&\frac 14(I_{4\times 4}+\sum_i a_i\sigma_i \otimes I_{2\times 2}+\sum_i b_i I_{2\times 2}\otimes \sigma_i\nonumber \\
&+&\sum_i c_i \sigma _i\otimes\sigma_i),
\end{eqnarray}
that is a density matrix completely defined by nine real parameters arranged in three $3$-dimensional column vectors $\vec{a}=\{a_i\}$, $\vec{b}=\{b_i\}$ and $\vec c =\{c_i\}$.
This follows from the fact that local unitary operations $\rho'=(U_A \otimes U_B) \rho (U_A \otimes U_B)^\dagger$ correspond to left and right multiplication of the Bloch matrix $R$ with orthogonal matrices \cite{verstraete},
\begin{equation}\label{ortho}
R'=\left(
\begin{array}{cc}
1 & 0 \\
0 & O_A^T \\
\end{array}
\right) R \left(
\begin{array}{cc}
1 & 0 \\
0 & O_B \\
\end{array}
\right)\,,
\end{equation}
with $O_{A,B} \in SO(3)$. It is then straightforward to obtain the normal form of Eq.~(\ref{simpler}): one needs to calculate the singular value decomposition of the lower diagonal $3\otimes3$ block $T$ of $R$, $T=O_A C O_B^T$, divide $O_A$ and $O_B$ by their respective determinant (to make sure they both have determinant $+1$), and then apply Eq.~(\ref{ortho}). The density matrix is correspondingly transformed into the normal form of Eq.~(\ref{simpler}), where the $c_i$ are identified with the elements of the diagonal matrix $C$, i.e., the singular values of $T$. Every two--qubit state can be then transformed in its simplified normal form by means of local unitaries (which preserve entanglement and correlations in general, by definition), so we can restrict our analysis to density matrices of this type without incurring in any loss of generality.
Now, we move to calculate the quantum discord ${\cal D}$, \eq{discord}, for generic states in normal form. The marginal entropy ${\cal S}(B)$ and the global entropy ${\cal S}(A,B)$ are trivial to obtain. The main issue regards the optimization involved in the conditional entropy.
\subsection{Conditional Entropy}
Firstly, we have to write the conditional entropy in an explicit form, adopting for simplicity the notation in \cite{alber}. We remarked that a measurement sends the state $\rho$ into an ensemble $\{p_k,\rho _k\}$ as expressed in \eq{measure}. The entropy of the ensemble can be written as
\begin{eqnarray}
\tilde {\cal S}&=&\sum_k p_k {\cal S}(A|B_{\{P_{Bk}\}})=\sum_k p_k {\cal S}(\rho_k)\\\nonumber
&=&p_0 S_0 +p_1 S_1,
\end{eqnarray}
where $S_0, S_1$ are the entropies associated to $\rho_0,\rho_1$.\\
The measurement is defined by the quantity $P_{Bk}$ and is consequently parametrized by the elements of the unitary matrix $V$, which we can write in the basis of the Pauli matrices as
\begin{eqnarray}
V&=&v_0 I_{2 \times 2} + \vec v \vec\sigma \\\nonumber
&=& \left(
\begin{array}{cc}
v_0+v_3 & v_1-iv_2 \\
v_1+iv_2 & v_0- v_3
\end{array}
\right).
\end{eqnarray}
We notice that the real vector $\{v_0,\vec v\} =\{v_i\}$ has norm one. Therefore, it is possible to rearrange the four parameters in three variables only, for example in this way,
\begin{eqnarray}
h&=& v_0v_1+v_2v_3\nonumber\\
j&=&v_1v_3-v_0v_2\\
k&=&v_0^2+v_3^2.\nonumber
\end{eqnarray}
Setting the vectors $\vec{X} =\{2 j,2 h, 2k-1\}$ and $\vec{m}_{\pm}=\{m_{i\pm}\}=\{a_i \pm c_i X_i\}$, we have that, after a straightforward calculation, the conditional entropy takes the following expression:
\begin{eqnarray}\label{scond}
\tilde {\cal S}\!&\!=\!&\!-\frac 14\left\{ (1-\vec{b}\cdot\vec{X})\Bigg[\left(1-\frac{| \vec{m}_-|}{1-\vec{b}\cdot\vec{X}}\right)\log_2\!\left(1-\frac{|\vec{m}_-|}{1-\vec{b}\cdot\vec{X}}\right)\right.\nonumber\\
&+&\left(1+\frac{|\vec{m}_-|}{1-\vec{b}\cdot\vec{X}}\right)\log_2\!\left(1+\frac{| \vec{m}_-|}{1-\vec{b}\cdot\vec{X}}\right)\Bigg]\nonumber\\
&+&(1+\vec{b}\cdot\vec{X})\Bigg[\left(1-\frac{|\vec{m}_+|}{1+\vec{b}\cdot\vec{X}}\right)\log_2\!\left(1-\frac{| \vec{m}_+|}{1+\vec{b}\cdot\vec{X}}\right)\nonumber \\
&+&\left.\left(1+\frac{| \vec{m}_+|}{1+\vec{b}\cdot\vec{X}}\right)\log_2\!\left(1+\frac{| \vec{m}_+|}{1+\vec{b}\cdot\vec{X}}\right)\Bigg]\right\}.
\end{eqnarray}
This result is consistent with the formula provided in the Appendix of \cite{cinesi_sensati}, but we have reached here a simpler expression by exploiting the normal form of the density matrix, \eq{simpler}.
However, we have to remark that in this picture there is still an amount of redundancy \cite{cinesi_sensati}. A projective measurement on a two--qubit state can be characterized by two independent variables only, identifiable as the angles $\theta$ and $\phi$, which parametrize a generic single--qubit pure state as $|\psi\rangle = \cos \theta |0\rangle +e^{i\phi}\sin\theta|1\rangle$, and the Bloch sphere of coordinates $\{x,y,z\}$ in this way,
\begin{eqnarray}
\left\{
\begin{array}{ccc}
x&=& 2j =2 \cos\theta\sin\theta\cos\phi\\
y&=& 2h=2\cos\theta\sin\theta\sin\phi\\
z&=& 2k-1=2 \cos ^2\theta-1.
\end{array}
\right.
\end{eqnarray}
It is immediate to verify that the following constraint holds
\begin{eqnarray}
k^2+h^2+j^2=k.
\end{eqnarray}
The algorithm originally designed for $X$--states in \cite{alber} is flawed just in not considering the mutual dependence of $h,j,k$, resulting reliable only for a more restricted class of states identified in \cite{nuovodiscord}.
The above mapping enables us to re-parameterize the conditional entropy, \eq{scond}, as a function of the azimuthal and polar angles $\theta,\phi$; we can then write $\tilde {\cal S}(h,j,k)=\tilde {\cal S}(\theta,\phi)$ and perform the optimization of $\tilde {\cal S}$ over these two independent variables.
\subsection{Optimization}
Inspired by \cite{alber}, we look for symmetries in the expression of the conditional entropy. We notice immediately the invariance under the transformation $\theta\rightarrow \theta \pm \pi$, which, however, can be englobed by the following one
\begin{eqnarray}
\left\{
\begin{array}{c}
k \rightarrow 1-k\\
h \rightarrow -h\\
j \rightarrow -j,\\
\end{array}
\right .
\end{eqnarray}
which corresponds to
\begin{eqnarray}
\theta \rightarrow \theta \pm \pi/2.
\end{eqnarray}
We can appreciate this with an example. Let us pick a random state, such as
\begin{equation}\label{random}
{\scriptsize
\left(\!
\begin{array}{cccc}
0.437 & 0.126 + 0.197 i & 0.0271 - 0.0258 i & -0.274 + 0.0997 i \\
0.126 -
0.197 i &
0.154 & -0.0115 - 0.0187 i & -0.0315 +
0.170 i\\
0.0271 + 0.0258 i & -0.0115 + 0.0187 i &
0.0370 &
0.00219 - 0.0367 i\\
-0.274 - 0.0997 i & -0.0315 -
0.170 i & 0.00219 + 0.0367 i & 0.372
\end{array}\!
\right)};
\end{equation}
we can operate with local unitaries on it, obtaining a new state $\rho$ (albeit with the same entropies and discord) described by the simplified normal form presented in \eq{simpler}; we then perform a projective measurement on subsystem $B$, obtaining an ensemble whose conditional entropy is plotted in Fig.~\ref{discordsample}.
\begin{figure}[tb]
\centering \includegraphics[width=8cm ]{sexample.eps}
\caption{(Color online) Example of conditional entropy $\tilde {\cal S}$ for a random two--qubit state [\eq{random}]. The angles $\theta$ and $\phi$ parametrize the measurement: we can appreciate the symmetry properties of $\tilde {\cal S}$ with respect to such variables, expressed by the invariance $\tilde {\cal S}(\theta,\phi)=\tilde {\cal S}(\theta \pm\pi/2, \phi)$. All the quantities plotted are dimensionless.}
\label{discordsample}
\end{figure}
One sees that there are no further apparent symmetries for the conditional entropy. Therefore the analysis so far, while not being conclusive, allows us just to refine the problem by safely letting the optimization of the conditional entropy to be restricted to the interval $\theta \in [0,\pi/2)$. To determine the minimum of $\tilde {\cal S}$, we need to calculate its derivatives with respect to $\theta$ and $\phi$. The dependence on these variables involves logarithms of non--linear quantities, so we can not expect to solve analytically the problem in any case, whatever ingenious variables we might choose. However, we can seek to write the two constraints in a compact and elegant form. Let us impose
\begin{eqnarray}
\left\{
\begin{array}{ccc}
p&=&\vec{b}\cdot\vec{X}\\
r_+&=&|\vec{m}_+ |\\
r_-&=&|\vec{m}_-|;
\end{array}
\right.
\end{eqnarray}
After a bit of algebra, we obtain
\begin{eqnarray}
\tilde {\cal S}&=&-\frac14\bigg\{(1 - p - r_-) \log_2[1 - p - r_-] \\\nonumber
&+& (1 - p + r_-) \log_2[ 1 - p + r_-] \\ \nonumber &+& (1 + p + r_+) \log_2[ 1 + p + r_+] \\\nonumber
&+&(1 + p - r_+) \log_2[ 1 + p - r_+] \\ \nonumber &-& 4 + 2 (-(1 - p) \log_2[ 1 - p] \\\nonumber
&-& (1 + p) \log_2[ 1 + p])\bigg\} .
\end{eqnarray}
Now, we set the partial derivatives to zero,
\begin{eqnarray}
\left\{
\begin{array}{c}
\frac{\partial \tilde {\cal S}}{\partial\theta}= \frac{\partial \tilde {\cal S}}{\partial p}\frac{\partial p}{\partial \theta}+\frac{\partial \tilde {\cal S}}{\partial r_+}\frac{\partial r_+}{\partial \theta}+ \frac{\partial \tilde {\cal S}}{\partial r_-}\frac{\partial r_-}{\partial \theta}=0\,;\\\nonumber \\ \nonumber
\frac{\partial \tilde {\cal S}}{\partial\phi}= \frac{\partial \tilde {\cal S}}{\partial p}\frac{\partial p}{\partial \phi}+\frac{\partial \tilde {\cal S}}{\partial r_+}\frac{\partial r_+}{\partial \phi}+ \frac{\partial \tilde {\cal S}}{\partial r_-}\frac{\partial r_-}{\partial \phi}=0\,.
\end{array}
\right .
\end{eqnarray}
Defining the following quantities
\begin{eqnarray}
\alpha&=&\det \left(
\begin{array}{cc}
\frac{\partial p}{\partial\theta} &\frac{\partial p}{\partial\phi}\\
\frac{\partial r_+}{\partial\theta}&\frac{\partial r_+}{\partial\phi}
\end{array},
\right),\nonumber\\
\beta&=&\det \left(
\begin{array}{cc}
\frac{\partial p}{\partial\theta} &\frac{\partial p}{\partial\phi}\\
\frac{\partial r_-}{\partial\theta}&\frac{\partial r_-}{\partial\phi}
\end{array}
\right),\\
\gamma&=&\det \left(
\begin{array}{cc}
\frac{\partial r_+}{\partial\theta} &\frac{\partial r_+}{\partial\phi}\\
\frac{\partial r_-}{\partial\theta}&\frac{\partial r_-}{\partial\phi}
\end{array}\nonumber
\right),
\end{eqnarray}
after some manipulations, we can write the stationarity conditions in the following form
\begin{eqnarray}
\left\{
\begin{array}{c}
\frac 14 \log_2 [\frac{1+p-r_+}{1+p+r_+}]+\frac 12 \log_2[\frac{(1+p)(1-p-r_-)}{(1-p)(1+p-r_+)}]\frac{\beta}{\alpha+\beta+\gamma}=0\,;\\ \\
\frac 14 \log_2 [\frac{1-p-r_-}{1-p+r_-}]-\frac 12 \log_2[\frac{(1+p)(1-p-r_-)}{(1-p)(1+p-r_+)}]\frac{\alpha}{\alpha+\beta+\gamma}=0\,.
\end{array}
\right .
\end{eqnarray}
We see immediately that this system can be further simplified to
\begin{eqnarray}
\left\{
\begin{array}{c}
\frac{\log_2 [\frac{1+p-r_+}{1+p+r_+}]}{\beta}+ \frac{\log_2 [\frac{1-p-r_-}{1-p+r_-}]}{\alpha}=0\,;\\ \\
\frac 14 \log_2 [\frac{1-p-r_-}{1-p+r_-}]-\frac 12 \log_2[\frac{(1+p)(1-p-r_-)}{(1-p)(1+p-r_+)}]\frac{\alpha}{\alpha+\beta+\gamma}=0\, .
\end{array}
\right .
\end{eqnarray}
We can still express these equations as relations among the eigenvalues of the ensemble $\{p_k,\rho_k\}$. Calling $\lambda_0^{+},\lambda_0^-$ the eigenvalues of $\rho_0$ and $\lambda_1^+,\lambda_1^-$ the eigenvalues of $\rho_1$, we have
\begin{equation}
\begin{split}
\lambda_0^{\pm}&=\frac12\left(1\pm\frac{r_-}{1-p}\right)\,,\\
\lambda_1^{\pm}&=\frac12\left(1\pm\frac{r_+}{1+p}\right)\,,
\end{split}
\end{equation}
After some straightforward algebra, one can show that the vanishing of the derivatives of $\tilde {\cal S}$ occurs when the following constraints are satisfied
\begin{eqnarray}\label{eletrash}
\left\{
\begin{array}{c}
\lambda_0^{-}=\frac{\left(\frac{\lambda_1^{+}}{\lambda_1^{-}}\right)^{\frac{\alpha}{\beta}}}
{1+\left(\frac{\lambda_1^{+}}{\lambda_1^{-}}\right)^{\frac{\alpha}{\beta}}}\,;\\ \\
\lambda_1^{-}= \lambda_0^{-}\left(\frac{\lambda_0^{+}}{\lambda_0^{-}}\right)^{\frac{\alpha+\beta+\gamma}{2\alpha}}\,.
\end{array}
\right .
\end{eqnarray}
These two transcendental equations can be solved numerically. They represent the most compact formulation to date for the problem of calculating the quantum discord of arbitrary two--qubit states. Let us call $s_i$ the solutions obtained, corresponding to values $\{\theta_i,\phi_i\}$. In order to establish if they represent minima of $\tilde {\cal S}$, we adopt the conventional method and evaluate the signature of the Hessian matrix $H$ at the points $\{\theta_i,\phi_i\}$, and, in case of $\text{det} H =0$, we study the sign of the functions $\delta_i=\tilde {\cal S}(\theta,\phi)-\tilde {\cal S}(\theta_i,\phi_i)$. Naming $\{\theta_{mj},\phi_{mj}\}$ the angles such that $H$ is positive definite or $\delta_i > 0$, we clearly have that the absolute minimum of the conditional entropy is defined as
\begin{equation*}
\text{min}_{\{P_{Bk}\}}\sum_k p_k {\cal S}(A|B_{\{P_{Bk}\}})= \text{min}_{\{\theta_{mj},\phi_{mj}\}}\tilde {\cal S}(\theta_{mj},\phi_{mj}).
\end{equation*}
The quantum discord for generic two--qubit states of the form \eq{simpler} finally reads
\begin{equation}\label{discordfinal}
{\cal D}(A:B)= {\cal S}(B)- {\cal S}(A,B) + \text{min}_{\{\theta_{mj},\phi_{mj}\}}\tilde {\cal S}(\theta_{mj},\phi_{mj}).
\end{equation}
\section{Comparison between quantum discord and geometric discord}\label{sec3}
We can use our results to compare the quantum discord ${\cal D}$ \cite{zurek} with the geometric discord $D_G$ \cite{dakic} for general two--qubit states \cite{comparezap}.
We have generated up to $10^6$ random general two--qubit states. After transforming each of them into the normal form of \eq{simpler}, we have calculated their quantum discord ${\cal D}$, as numerically obtained from the algorithm of Section \ref{sec2}, and their normalized geometric discord $2 D_G$. The latter admits the following explicit analytic expression for states $\rho$ in normal form, derived from \eq{dgformula},
\begin{eqnarray}
D_G (\rho)= \frac 14(||\vec b \vec b^T||_2 + ||\vec c \vec c^T||_2 -\tilde k),
\end{eqnarray}
where $\tilde k$ is the largest eigenvalue of the matrix $\vec b \vec b^T + \vec c \vec c^T$.
The results are shown in Fig.~\ref{zurVSgeo}. We notice that physical states of two qubits fill up a two-dimensional area in the space of the two non-classicality measures, meaning that the two impose inequivalent orderings on the set of mixed two--qubit quantum states; this is reminiscent of the case of entanglement measures (see e.g. \cite{nemoto}), and a similar feature has been reported concerning the comparison between discord and other non-classicality indicators \cite{gap}. Nevertheless, at fixed quantum discord, the geometric discord admits exact lower and upper bounds (and vice versa). We have identified them numerically.
\begin{figure}[tb]
\centering
\includegraphics[width=8.5cm ]{zurVSgeo.eps}
\caption{(Color online) Comparison between normalized geometric discord $2D_G$ and quantum discord ${\cal D}$ for $10^6$ randomly generated general two--qubit states. The dashed line is obtained by taking the equality sign in \ineq{hier}. Refer to the main text for details of the other boundary curves. All the quantities plotted are dimensionless.}
\label{zurVSgeo}
\end{figure}
Firsts of all, we have observed that the following hierarchical relationship holds for arbitrary two--qubit states,
\begin{equation}\label{hier}
2D_G \ge {\cal D}^2\,.
\end{equation}
In other words, the quantum discord for any two--qubit state can never exceed the square root of its (normalized) distance from the set of classical--quantum states. The corresponding boundary curve is plotted as a dashed (gray) line in Fig.~\ref{zurVSgeo}. However, we see that such a bound is tight only in the region $0 \le {\cal D} \le 1/3$, in which it coincides with what we refer to as branch (i) [see below], while it is not attainable for higher degrees of non-classical correlations. The actual, tight lower bound in the whole $\{{\cal D}, 2D_G\}$ plane accommodates states with minimal geometric discord at fixed quantum discord, or equivalently maximal quantum discord at fixed geometric discord: Such extremal states are constituted by the union of four different families, which sit on branches (i)--(iv) in Fig.~\ref{zurVSgeo}.
\begin{enumerate}
\item[(i)] (Green online) This branch is filled by so-called $\alpha$ states \cite{james}
\begin{equation}\label{alphastates}
\rho_\alpha= \left(
\begin{array}{cccc}
\frac{\alpha }{2} & 0 & 0 & \frac{\alpha }{2} \\
0 & \frac{1-\alpha }{2} & 0 & 0 \\
0 & 0 & \frac{1-\alpha }{2} & 0 \\
\frac{\alpha }{2} & 0 & 0 & \frac{\alpha }{2}
\end{array}
\right)\,,\quad 0 \le \alpha \le 1/3\,,
\end{equation}
for which
${\cal D}(\rho_\alpha) = \alpha$ and $2D_G(\rho_\alpha) = \alpha^2$,
thus saturating \ineq{hier}.
\item[(ii)] (Blue online)
This small branch is filled by a subclass of the two-parameter family
\begin{equation}\label{arstates}\begin{split}
& \rho_{r}=\left(
\begin{array}{cccc}
(1-a)/2 & 0 & 0 & r/2 \\
0 & a & 0 & 0 \\
0 & 0 & 0 & 0 \\
r/2 & 0 & 0 & (1-a)/2
\end{array}
\right)\,, \\ &\frac13 \le a \le \frac{5}{14}\,,\quad \sqrt{4a-3a^2-1}\,,\end{split}\end{equation}
with $r \in \left[ \sqrt{4 a-3 a^2-1},\, \frac{1-a}{3}\right]$ given by the solution to $\frac{2 r \tanh ^{-1}\left(\sqrt{a^2+r^2}\right)}{\sqrt{a^2+r^2}}+\ln (-a-r+1)-\ln (-a+r+1)+2 \tanh ^{-1}(r)=0$. The geometric discord of these states is simply $2D_G(\rho_{r})=a^2$, while their quantum discord is calculated in \cite{james}.
We highlight the presence of the ``pimple'' at the joint between branches (i) and (ii), a recurring feature in the profile of extremal states involving quantum discord \cite{james,galve,gap}.
\item[(iii)] (Red online) This branch accommodates asymmetric $X$--states of the form
\begin{equation}\label{asymx}\begin{split}
\rho_g=\left(
\begin{array}{cccc}
a & 0 & 0 & \sqrt{a-a^2-a c} \\
0 & c & 0 & 0 \\
0 & 0 & 0 & 0 \\
\sqrt{a-a^2-a c} & 0 & 0 & 1-a-c
\end{array}
\right)\,, \\ a = \frac{1-2 c+2 c^2-g}{2 c}\,, \,\,0\le g \le 1\,,\end{split}
\end{equation}
with $$c \in \left[\frac{1-\sqrt{g}}2,\, \frac12 - \left\{
\begin{array}{ll}
\frac12\sqrt{2g-1}, & g>\frac12; \\
0, & \hbox{otherwise;}
\end{array}
\right.\right]\,,
$$
solution to $8 (1-2 c) c^2 \tanh ^{-1}\left(\sqrt{8 (c-1) c-2 g+3}\right)-4 c^2 \sqrt{8 (c-1) c-2 g+3} \tanh ^{-1}(1-2 c)+2 \sqrt{8(c-c^2)-2 g+3} \left(2 c^2+g-1\right) \tanh ^{-1}\left(\frac{3c-2 c^2 +g-1}{c}\right)=0$. For these states, $2D_G(\rho_g)=g$ and ${\cal D}(\rho_g) = \frac{1}{\ln 4}\, \bigg[-\ln (-4 c (a+c-1))-2 \sqrt{4 c (a+c-1)+1} \tanh ^{-1}\left(\sqrt{4 c (a+c-1)+1}\right)-2 \ln (1-a)+4 a \tanh ^{-1}(1-2 a)+2 \ln (2-2 c)-4 c \tanh ^{-1}(1-2 c)\bigg]$.
\item[(iv)] (Black online) The top-right-most branch accommodates just pure states $\rho_p=\ket{\psi}_{AB}\bra{\psi}$, for which the discord equals the marginal von Neumann entropy, ${\cal D}(\rho_p)={\cal S}(\rho_A)=-p \log_2 p - (1-p) \log_2 (1-p)$, and the geometric discord equals the marginal linear entropy, $2D_G(\rho_p) = 2(1-\text{Tr} \rho_A^2) = 4 \det \rho_A=4p(1-p)$, where we have denoted the eigenvalues of the reduced density matrix $\rho_A$ by $\{p, 1-p\}$.
\end{enumerate}
\smallskip
\noindent
On the other hand, the upper boundary (v) in Fig.~\ref{zurVSgeo}, despite being single--branched, is more involved and we are unable to provide a tractable parametrization of the states that saturate it. They can be sought among symmetric $X$--states of full rank, but with the two biggest eigenvalues dominating the other two. The extremal curve has been obtained as the result of extensive numerical optimization, in which the parameter space has been finely sliced in discrete intervals of nearly constant discord, and for each interval the datapoint corresponding to the random state with the maximum geometric discord has been selected. Joining all such extremal states we have obtained the smooth (Magenta online) line of Fig.~\ref{zurVSgeo}.
The two measures ${\cal D}$ and $2D_G$ correctly coincide on classical--quantum states \eq{cq}, where both vanish, and on maximally entangled Bell states, where both reach unity.
\section{Conclusions}\label{sec4}
We have presented a reliable and effective algorithm for the evaluation of the quantum discord ${\cal D}$ of general two--qubit states. We have simplified the optimization involved in calculating the conditional entropy, by removing the redundant degrees of freedom that can be set to zero by means of local unitaries in the first place, and by properly taking into account the symmetries of the problem. The optimization problem for the conditional entropy, and equivalently for the discord, is recast into a compact form that implies an elegant relationship among the eigenvalues of the ensemble obtained after the local measurement process on one qubit. The derived transcendental constraints are amenable to direct numerical solution.
We have then compared quantum discord with an alternative but affine quantity, the geometric discord $D_G$, identifying the classes of states with extremal values of geometric discord at fixed quantum discord. For a fixed geometric discord, maximal quantum discord is attained by different families of states depending on the degree of non-classical correlations, encompassing pure as well as mixed, symmetric and nonsymmetric states. In general, the hierarchical bound ${\cal D} \le \sqrt{2 D_G}$ holds for all two--qubit states.
We hope that our results can provide further insight into the fascinating but still not well-understood paradigm of non-classical correlations beyond entanglement in composite quantum systems. The methods presented here can be generalized to higher--dimensional and continuous variable systems.
\acknowledgments{We thank Mauro Paternostro and Steve Campbell for very fruitful exchanges during the early stages of this project.}
|
\section{Introduction}
\label{sec:Intro}
Automated demand response is often used to manage electrical load during critical system peaks\cite{07LBNL-DR-report,DOE-DR06}. During a typical event as the system approaches peak load, signaling from the utility results in automated customer load curtailment for a given period of time to avoid overstressing the grid. Although this type of load control is useful for maintaining system security, automated demand response must evolve further to meet the coming challenge of integrating time-intermittent renewables such as wind or photovoltaic generation. When these resources achieve high penetration and their temporal fluctuations exceed a level that can be economically mitigated by the remaining flexible traditional generation (e.g. combustion gas turbines), automated demand response will play a large role in maintaining the balance between generation and load. To fill this role, automated demand response must go beyond today's peak-shaving capability
To follow intermittent generation, automated demand response must be bi-directional control, i.e. it should provide for controlled increases and decreases in load. The response must also be predictable and preferably non-hysteretic, otherwise the load-generation imbalance may actually be exacerbated. Predictability would be highly valued by third party companies that aggregate loads into a pool of demand response resources. Finally, whatever control methodology is implemented, it must also be stable and not exhibit temporal oscillations. There are several factors that make achieving these demand response goals challenging: the different options for demand response signal, the uncertainty of the aggregate response to that signal, and the inhomogeneity of the underlying ensemble of loads.
The demand response control signal could take several forms: direct load control where some number of loads could be disabled via a utility-controlled switch\cite{92SS,06BKT}; end-use parameter control where an ensemble of loads can be controlled by modifying the set point of the end-use controller, e.g. a thermostat temperature set point\cite{09Cal,11CH,10KSBH}; or indirect control via energy pricing in either a price taking (open loop) or auction (closed loop)\cite{Olympic} setting. Today's automated demand response for peak-shaving is a form of direct load control which could be adapted and refined for the type of operation we desire, however, it is difficult to assess the impact of demand response on the end user because loads are simply disabled and re-enabled with little concern for the current state of the end use. Direct control is feasible for a relatively small number of large loads because the communication overhead is not extreme. Individual direct control of a large number of small loads would potentially overburden a communication system, however, ``ensemble'' control using a single parameter for control has been proposed, e.g. set point control for thermostatic loads\cite{09Cal,11CH,10KSBH} and connection rate control for electric vehicle charging\cite{10TSBC}. However, in these control models, the underlying loads are assumed to be homogeneous (all of the same type), which is advantageous because it allows for a quantifiable measure of the end use impacts and customer discomfort, e.g. increasing all cooling
thermostat set points by 1$^oF$ will generate a decrease in load with a known end-use impact.
To control a large ensemble of {\it inhomogenous} loads with a single demand response signal requires a quantity that applies to all loads, i.e. energy pricing \cite{80STKOPC}. When given access to energy prices, consumers (or automated controllers acting on their behalf) can make their own local decisions about whether to consume or not. These local decisions open up new possibilities and also create problems. The customer is now enabled to automatically modify and perhaps optimize his consumption of energy to maximize his own welfare, which is a combination of his total energy costs and the completion of the load's end use function. However, without an understanding of how consumers respond to energy prices, the fidelity of the control allowed by the direct or ensemble control schemes described above is lost. Retail-level double auction markets\cite{Olympic} are an effective way of making demand response via pricing a closed-loop control system, however, a logical outcome of these markets would be locational prices potentially driven distribution system constraints making the regulatory implementation troublesome. In contrast, a model where retail customers are price takers may avoid some regulatory issues, however, price taking is in essence a form of open loop control which then requires an understanding of how the aggregate load on the system will respond to price.
Our goal in this initial work is to layout the computational framework for discovering the end-use response to these price-taking ``open-loop'' control systems. We develop state models for several different loads and subject them to a stochastic price signal that represents how energy prices might behave in an grid with a large amount of time-intermittent generation. We analyze the response of these smart loads using a Markov Decision Process (MDP) to optimize the welfare of the end user. Human owners of the devices have the ability to program the devices in accordance to their strategies and preferences, for instance by adjusting their willingness to sacrifice comfort in exchange for savings on electricity costs. Otherwise, most of the time we assume that the devices operate automatically in accordance to some optimal algorithm that was either preprogrammed by their owners, discovered via adaptive learning\cite{10OLGM}, or programmed by a third-party aggregator. The resulting load end-use policies can then be turned around to predict the effect of a change in prices on electrical load. Our long term strategic intention is to analyze the aggregated network effect on power flows of many independent customers and design optimal strategies for both consumers and the power operator. However, the prime focus of our first publication on the subject is less ambitious. We focus here on description of different load models and analyze the optimal behavior of individual consumers.
The material in the manuscript is organized as follows. We formulate our main assumptions and introduce the general MDP framework in Section \ref{sec:setting}. Models of four different devices (optional, deferable and control loads and storage devices) are introduced in
Section \ref{sec:devices}. Our enabling simulation example of a control load (smart thermostat) is presented in Section \ref{sec:Simulations}. We summarize our main results and discuss a path forward in Section \ref{sec:Path}.
\section{Setting the Problems}
\label{sec:setting}
\subsection{Basic Assumptions}
\label{subsec:Basic Assumptions}
Future distribution networks are expected to show complex, collective behavior originating from competitive interaction of individual players of the following three types:
\begin{itemize}
\item Market operator, having full or partial control over the signals sent to devices/customers. The most direct signal is energy price. The operator may also provide subsidies and incentives or impose penalties, however in this manuscript, we will mainly focus on direct price control
\item Human customers/owners, who are able to reprogram smart-devices or override their actions.
\item Smart devices, capable of making decisions about their operations. The devices are semi-automatic, i.e. pre-programmed to respond to the signal on a short time scale (measured in seconds-to-minutes) in a specific way, however the owner of the device may also choose to change the strategy on a longer time-scale (days or weeks). We model the smart devices as finite state machines using a Markov Decision Process (MDP) framework. At the beginning of each interval, a device decides how to change its state based on the current price. Each change comes with a reward expressing actual transactions between the provider and the consumer and the level of consumer satisfaction with the decision. We assume that smart devices are selfish and not collaborative, each optimizing its own reward.
\end{itemize}
In this manuscript we restrict our attention to a simple price-taking strategy of consumer behavior, deferring analysis of more elaborate game-theoretic interactions between the operator and the individual customers to further publications.
We model the external states (that include electricity price, weather, and human behavior) as a stochastic, Markov Chain process, $\{s^{(e)}(t)\}$. At the beginning of the time interval, $t$, the variable describing these factors is set to $s^{(e)}_t$ and changes during the next time step to $s^{(e)}_{t+1}$ with the transition probability $T(s^{(e)}_{t+1}|s^{(e)}_t)$. The transition probabilities are assumed to be known to the device and statistically stationary, i.e. independent of $t$. (The later assumption can be easily relaxed to account for natural cycles and various external factors.) The probability, $p(s^{(e)};t)$, to observe the external state, $s^{(e)}(t)=s^{(e)}$, at the time $t$, follows the standard Markov chain equation
\begin{eqnarray} \label{external-chain}
p(s^{(e)};t+1)=\sum_{s_{t}^{(e)}} T(s^{(e)},s_{t}^{(e)}) p(s_t^{(e)}|t).
\label{P-eq}
\end{eqnarray}
We also assume that the Markov chain (\ref{external-chain}) is ergodic and converges after a finite transient to the statistically stationary distribution: $p(s^{(e)};t+1)=p(s^{(e)};t)=p(s^{(e)})$. In the simulation tests that follow we will restrict ourselves to $s^{(e)}$ drawn from a finite set $S^{(e)}$.
\subsection{General Markov Decision Process Framework}
\label{subsec:FSM}
Here we adopt the standard (Markov Decision Process) MDP approach \cite{57Bel,05Put,MDP-package} to the problem of interest: description of smart devices responding to the external (exogenous) Markov process $\{s^{(e)}(t)\}$. MDPs provide a mathematical framework for modeling decision-making in situations where outcomes are partly random and partly under the control of a decision maker. Formally, the MDP is a 4-tuple, $(S,A,P(\cdot,\cdot),R(\cdot,\cdot))$, where
\begin{itemize}
\item $S$ is the finite set of states, in our case a direct product of the machine states set $S^{(m)}$, and the externality state set $S^{(e)}$, $S=S^{(m)}\otimes S^{(e)}$.
\item $A$ is a finite set of actions. $A_s$ is the finite set of actions available from state $s\in S$. Within our framework we model only the decisions made by the machine, so the set $A$ consists only of actions associated with the machine, $A = A^{(m)}$.
\item $P_a(s,s')= \Pr(s_{t+1}=s' \mid s_t = s,\, a_t=a)$ is the probability that action $a$ chosen while in state $s=(s^{(m)},s^{(e)})$ at time $t$ will lead to state $s'$ at time $t + 1$. The probabilistic description of the transition allows to account for stochastic nature of the price fluctuations as well as for the randomness in the dynamics of the smart devices.
\item $R_a(s,s')$ is the reward associated with the transition $s \to s'$ if the action $a$ was chosen. In our models, the reward will reflect the price paid for electricity consumption associated with the transition as well as the level of discomfort related to the event.
\end{itemize}
In the most simple setting analyzed in this work, the behavior of the device is modeled via the policy function $\pi(s) : S\to A$ that determines the action chosen by the device for a given state: $a_t=\pi(s_t)$. More general formulations that include randomized decision making process, are not considered in this paper. Our smart device models seek to operate with the policy, $\pi(s)$, that maximizes over actions the expectation value of the total discounted reward, $\sum_{t=0}^\infty \gamma^t R_{a_t}(s_t,s_{t+1})$ over the Markov process, $P_a(\cdot,\cdot)$, where $0<\gamma\leq 1$, is the discount rate. There are numerous algorithms used for optimizing the policies. In our work we use the algorithms implemented in MDP Matlab toolbox \cite{MDP-package}.
\section{Models of Devices}
\label{sec:devices}
The specifics of our MDP setting are to be described below for four examples of loads. Note that these examples are meant to illustrate the power of the framework and its applicability to "smart grid" problems. In this first paper, we do not aim to make the examples realistic. Instead, we focus on the qualitative features of the loads
The states and actions associated with the devices are illustrated in the diagrams shown in Figs. \ref{fig:optional}-\ref{fig:storage}. For simplicity, we ignore the external part of the state $s^{(e)}$ in these diagrams. Full diagrams can be produced by taking the Kronecker product of transition graphs associated with the device and the external factors. In our diagrams, the states are marked by squares and actions are marked by dashed circles. Transitions from states to actions and actions to states are marked by dashed and solid arrows, respectively.
\begin{figure}[b]
\begin{center}
\includegraphics[width=0.45\textwidth]{optional.pdf}
\caption{MDP diagram for the model of optional load. See text for explanations.}
\label{fig:optional}
\end{center}
\end{figure}
\subsection{Optional Loads}
\label{subsec:OptLoads}
A smart device described by an ``optional load'' pattern can operate in two regimes, at full and limited capacity. An example of such load is a light that can be automatically dimmed if the electricity price becomes too high (see Fig.~\ref{fig:optional}). To simplify the mathematical notations, we denote the states of the machine $s^{(m)}$ by $x$. The machine can be in either of the two states: $x=0$ and $x=1$, shown as $Idle$ and $Active$ in the diagram (\ref{fig:optional}) respectively. In the $x=0$ state the machine does not operate (the lights are off). In the $x=1$ state, the machine is active and the lights are shining at the full brightness, or are dimmed. Actions of the device are $a_0 = \mbox{pass}$, $a_1=\mbox{full}$ or $a_2=\mbox{shed}$. The $a_0$ action represents the process of waiting for the external signal of switching on the device. If no external external signal (requesting switching on) appears, the system returns to the $x=0$ state, otherwise it moves to the $x=1$ state. When the device is active (in the $x=1$ state), it has two options: operate at full capacity, corresponding to the action $a_1$, or shed the load (dim the lights), corresponding to action $a_2$. Turning the device on or off is an externality dependent on a human. We assume that the external/human action is random, with the probability of turning the device ON and turning the device OFF being $\rho_{ON}$ and $\rho_{OFF}$ respectively. (For simplicity, we assume that the OFF signal may arrive only by the end of the time interval.) Assuming additionally that the transition probabilities do not not depend on the device actions, we arrive to the following expression for the transition kernel:
\begin{eqnarray} \label{transition}
& P_{pass}(s,s') = T(c'|c)\left[\rho_{ON} \delta_{x',1} + (1-\rho_{ON})\delta_{x',0}\right], \\
& P_{full,shed}(s,s') = T(c'|c)\left[\rho_{OFF} \delta_{x',0} + (1-\rho_{OFF})\delta_{x',1}\right],
\end{eqnarray}
where $\delta_{x_1,x_2}$ is the Kronecker symbol: it is unity if $x_1=x_2$ and zero otherwise.
There is no reward associated with either outcome of the $a_0 = \mbox{pass}$ action,
however, the other two actions ($a_1$ and $a_2$) result in a reward consisting of two contributions. First is the price paid to the
electricity provider, $E_{full,shed} c$, where $c(t)$ is the cost of electricity (considered as a component of $s^{(e)}$) and $E_{full,shed}$ is the amount of energy consumed during the time interval which depends on whether the lights are fully on or dimmed. ( Here, $E_{full}>E_{shed}>0$ and both values do not depend on the resulting state of the device). Second, the reward function accounts for a subjective level of comfort associated with the $a_{1,2}$ actions: $C_{full,shed}$. The discomfort of the light dimming is accounted by choosing $C_{full} > C_{shed}$. Summarizing, the cumulative reward function in this model of the optional load becomes
\begin{eqnarray}
& R_{pass}(s,s') = 0, \\
& R_{full}(s,s') = C_{full} - E_{full} c, \\
& R_{shed}(s,s') = C_{shed} - E_{shed} c.
\end{eqnarray}
Obviously, our model of optional loads is an oversimplification because there are a variety of additional effects which may also be important in practice, however, all these can be readily expressed within the MDP framework. For example, one may need to limit the wear and tear on the device, thus encouraging (via a proper reward) minimization of switching. (To account for this effect would require splitting the $Active$ state in the model explained above into two states $Active-Full$ and $Active-Shed$.)
\subsection{Deferable Loads}
\label{subsec:DefLoads}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{deferable.pdf}
\caption{MDP diagram for the model of deferable load. See text for explanations.}
\label{fig:deferable}
\end{center}
\end{figure}
Our second example model is a deferable load, i.e. a load whose operation can be delayed without causing a major consumer discomfort. Practical examples include dishwashing machines or some maintenance jobs like disk defragmentation on a computer. A simple model of such a device, shown in Fig. \ref{fig:deferable}, has two states: $x=0$ ($Idle$) when no work is required and $x=1$ ($Waiting$) when a job has been requested and the machine is waiting for the right moment (optimal in terms of the cost) to execute it. As in the previous model, the only action of the machine in the $Idle$ state is $a_0$ ($Pass$), however, in the $Waiting$ state, there are two possible actions: $a_1=Wait$ results in waiting for possible drop of the electricity price and $a_2=Work$ results in immediate execution of the job. The transition kernel for the model is
\begin{eqnarray}
& P_{Pass}(s,s') = T(c'|c)\left[\rho_{ON} \delta_{x',1} + (1-\rho_{ON})\delta_{x',0}\right], \\
& P_{Wait}(s,s') = T(c'|c)\delta_{x',1}, \\
& P_{Work}(s,s') = T(c'|c)\delta_{x',0},
\end{eqnarray}
where $\rho_{ON}$ is the probability of an exogeneous job request. In this model, there is no reward for choosing the $a_0=Pass$ action. The reward for the $a_2=Work$ action is equal to minus the price paid for the electricity, $R_{Work}(s,s') = - E*c$, and the reward for the $a_1=Wait$ action represents the level of discomfort associated with the delay, $R_{Wait} = C_{delay} < 0$. As in the model of optional loads, $E$ and $C_{delay}$ are constants parameters.
\subsection{Control Loads}
\label{subsec:ContrLoads}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{control.pdf}
\caption{MDP diagram for the model of controllable load. See text for explanations.}
\label{fig:control}
\end{center}
\end{figure}
A very important class of devices that will likely play a key role in future demand response technologies are machines tasked to maintain a prescribed level of physical characteristics of some system. For example, thermostats are tasked with keeping the temperature in a building within acceptable bounds. Other examples of the control devices are water heaters, electric ovens, ventilation systems, CPU coolers etc.
In our enabling, proof-of-principle model of the control load, we consider a thermostat responsible for temperature control in a residential home. The state of the device is fully characterized by temperature which can take three possible values: $x=0,1,2$ corresponding to $Low, Medium, High$ temperatures, respectively. Each temperature is assumed to be operationally acceptable. For simplicity, we assume that the thermostat uses an electric heater to modify the temperature (i.e. the outside temperature is low). The device can choose between the following three actions. $a_0 = Cool$ leaves the heater idle for the forthcoming interval. Since there is some base consumption associated with the thermostat operation we assume that $E_{Cool}>0$. The next action, $a_1=Keep$, maintains the temperature at the current level and requires some energy for heater operation: $E_{Keep}>E_{cool}>0$. Finally, $a_2=Heat$ corresponds to intensive heating that raises the temperature and requires the largest amount of energy $E_{Heat}$, and $E_{Heat}>E_{Keep}>E_{Cool}=0$. Our thermostat state diagram, shown in Fig.~(\ref{fig:control}), assumes that the dynamics of the thermostat are deterministic, and the resulting state depends only on the action chosen. The transition probabilities of the thermostat MDP is
\begin{eqnarray}
& P_{Heat}(s,s') = T(c'|c)\delta_{x',x+1}, \\
& P_{Keep}(s,s') = T(c'|c)\delta_{x',x}, \\
& P_{Cool}(s,s') = T(c'|c)\delta_{x',x-1}.
\end{eqnarray}
Assuming that all levels of temperature are equally comfortable, the reward function depends only on the price and energy consumption associated with the action,
\begin{eqnarray}
R_{Cool,Keep,Heat}(s,s') = - c E_{Cool,Keep,Heat}.
\label{reward_control}
\end{eqnarray}
Our basic model can be generalized to account for different comfort levels of different states, the possibility for the owner to override an action, variations of the outside temperature, etc.
\subsection{Storage loads}
\label{subsec:Storage}
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{storage.pdf}
\caption{MDP diagram for the model of storage. See text for explanations.}
\label{fig:storage}
\end{center}
\end{figure}
The number of devices with rechargeable batteries is expected to increase dramatically in the coming years. Currently, these are mostly laptops, uninterruptable power supplies, etc. In addition, a significant number of large-scale batteries will be added to the grid most likely via the anticipated Plug-in Hybrid Electric Vehicles (PHEV) potentially enabled with Vehicle-to-Grid (V2G) capability. Storage devices, illustrated with the MDP in Fig.~(\ref{fig:storage}), share some similarity with the controlled loads discussed in the previous Subsection, but they are also different in two aspects. First, users/owners wants their devices to be charged which leads to a level of discomfort if the devices are not fully charged. Second, and probably most significantly, storage devices such as PHEVs are disconnected from the grid when in use. Having PHEVs in mind, we propose the following model of (mobile) storage. The system can be in either of the three states, the $x=0=Unplugged$ state (which is similar to the Idle state in the models of Optional and Deferable loads discussed above), the $x=1=Partially$ state where the storage is partially charged, and the $x=2=Full$ state where the device is fully charged.
The four available actions are: $a_0=Pass$ when the device is in the unplugged state, the $a_1 = Keep$ action possible when the initial state is $x=1=Partially$ or $x=2=Full$, the $a_2=Charge$ action available from the $x=1=Partially$ state which transitions to the $x=2=Full$ state, and, finally, the $a_3=Discharge$ action, that is an inverse of the $a_2$ one, available from the $x=2=Full$ state resulting in the $x=1=State$. Except for $a_0=Pass$, all these actions can be interrupted by transitioning at the end of the time interval to the $x=0=Unplugged$ state. As in previous sections, we assume that the unplugging happens at the end of a time interval. Assuming the device can be unplugged with the probability $\rho_{OFF}$ and that it can be reconnected to the grid with the probability $\rho_{ON}$, we arrive at the following expressions for the transition probability:
\begin{eqnarray}
& P_{Pass}(s,s') = T(c'|c)\left[\rho_{ON}\delta_{x',1} + (1-\rho_{ON})\delta_{x',0}\right], \\
& P_{Keep}(s,s') = T(c'|c)\left[\rho_{OFF}\delta_{x',0} + (1-\rho_{OFF})\delta_{x,x'}\right], \\
& P_{Charge}(s,s') = T(c'|c)\left[\rho_{OFF}\delta_{x',0} + (1-\rho_{OFF})\delta_{x',2}\right],\\
& P_{Discharge}(s,s') = T(c'|c)\left[\rho_{OFF}\delta_{x',0} + (1-\rho_{OFF})\delta_{x',1}\right].
\end{eqnarray}
The reward function accounts for the following effects. First, the $a_1=Keep$ action has the cost associated with keeping the battery charged, $E_{Keep}(x)$, naturally dependent on the state, $E_{Keep}(2)>E_{Keep}(1)>E_{Keep}(0)=0$. Second, the $a_2=Charge$ action requires $E_{Charge}$ of energy while the $a_3=Discharge$ action generates the $E_{Discharge} < 0$ of energy, both nonzero only if the resulting state is not the $x=0=Unplugged$. Therefore, all the ``active" actions, $Keep, Charge, Discharge$, contribute the reward function in accordance with the energy price, $c' E_{\dots}$. Finally, we also assign an additional negative reward, $C_{Unplug} <0 $, accounting for the discomfort (to the human) associated with being in the $x=0=Unplugged$ state. The resulting reward function is
\begin{eqnarray}
& R_{Pass}(s,s') = 0,\\
& R_{Keep}(s,s') = C_{Unplug} \delta_{x',0}\delta_{x,1} - c E_{Keep}(x),\\
& R_{Charge}(s,s') =-c E_{Charge}, \\
& R_{Discharge}(s,s') = C_{Unplug}\delta_{x',0} - c E_{Discharge} .
\end{eqnarray}
\section{Simulations}
\label{sec:Simulations}
In order to illustrate the capabilities of the proposed framework, we consider a simple model of the control load, describing a smart thermostat, characterized by $N_T=10$ levels of the temperature parameter $T$. At every moment of time the thermostat can choose to raise, lower or keep the same temperature. The raise and lower options are not available at the highest and lowest possible temperatures, respectively. The energy consumption associated with the actions is given by $E_{Keep} = 1.0$, $E_{Cool} = 0.1$ and $E_{Heat} = 2.1$, respectively, in some normalized energy units. This choice of energies discourages the system from switching the heater too often: although the combinations $Heat + Cool$ and $Keep + Keep$ lead to the same temperature levels, the latter action is preferable as it consumes less energy.
Variations in price are modeled by a Markov chain of $N_P=5$ equidistant levels with the minimum and maximum corresponding to $1.0$ and $2.0$ price units, respectively. At each time interval, the price either increases with probability $T(c+1|c) = 0.5$ by $1$ level, decreases with probability $T(c-1|c) = 0.3$ by $1$ level, or stays the same. The resulting stationary probability distribution $p(c)$ is shown in the Figure \ref{fig:price-dist}. It is skewed towards the higher price, mimicking the effect of intermittent renewable generators that occasionally provide excess power to the grid, thus leading to rapid dips in the price.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{price-dist.pdf}
\caption{Probability distribution of electricity price in the model example.}
\label{fig:price-dist}
\end{center}
\end{figure}
The reward function (\ref{reward_control}) is fully determined by the total cost of energy consumed by the thermostat within the given time-interval. Our MDP model imposes upper and lower bounds on the temperature, and we assume that there is no additional discomfort associated with the variations of temperature between these bounds, i.e. all of the $N_T$ temperature levels are equally comfortable for the consumer.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{policy.pdf}
\caption{Visualization of the policy found as a result of optimization.}
\label{fig:policy}
\end{center}
\end{figure}
This system was analyzed with the Matlab MDP package \cite{MDP-package} where we used different algorithms to verify the stability of the results. The resulting optimal policy (for the range of parameters tested) is illustrated in \ref{fig:policy}. As expected, the thermostat chooses the $Heat$ action when the price is low and decides to $Cool$ when the price is high; a set of actions that lead to the skewed probability distribution of temperatures shown in Figure \ref{fig:temperature}. One finds that the thermostat spends most of the time performing $Keep$ in the low temperature state waiting for the price to drop.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{temperature.pdf}
\caption{Probability distribution of temperature levels observed at the optimal policy.}
\label{fig:temperature}
\end{center}
\end{figure}
Perhaps, the most interesting feature of the MDP model is the relation between consumption and price. We define the expected demand as the average energy demand for a given price
\begin{equation}
\langle E|c\rangle = \frac{\sum_{x} E_{\pi(x,c)} P_{st}(x,c) }{\sum_{x} P_{st}(x,c)},
\label{elacticity}
\end{equation}
where $P_{st}(x,c)$ is the stationary joint distribution function of the temperature and price at the optimal strategy. Dependence of the consumption on the price for our choice of the parameters is shown in Figure \ref{fig:elasticity}, thus illustrating that variations in price indeed produce demand response. An interesting feature is that the demand curve is not monotonic. At low temperatures, the energy consumption shows a slight increase with the price; a surprising behavior related to saturation of the demand. When the electricity price decreases gradually from high to low levels, there is a high probability that the thermostat will reach the highest level of temperature before the price reaches the lowest level. In this case, the demand will be lower at the smallest price levels as there will be no unsatisfied demand left in the system to capitalize on the lowest price. From the economic viewpoint, it is important to note that this non-monotonicity of the demand curve reflects the adaptive nature of the MDP algorithm: the smart devices adjust to fluctuations in price, thus making it more difficult for the electricity providers to exploit the non-monotonic demand curve for making profit.
\begin{figure}
\begin{center}
\includegraphics[width=0.45\textwidth]{elasticity.pdf}
\caption{Demand of the smart thermostats.}
\label{fig:elasticity}
\end{center}
\end{figure}
Another interesting result found in our simulations is an increase in average consumption of the smart (policy optimized) thermostat when compared to its non-smart counterpart, where the latter is defined as the one ignoring price fluctuations and sticking to the $Keep$ action. For the set of parameters chosen in the test case, we observed that the average level of consumption in the optimal case is $1.03$, i.e. it is $3\%$ higher than in the naive strategy, an effect associated with the additional penalty (in energy) imposed on the $Heat$ and $Cool$ actions.
It is also instructive to evaluate savings of the consumer. The average value of the reward associated with the optimal policy is equal to $-1.6722$, which should be compared with the reward of $-1.73$ generated by its non-smart counterpart. Since the reward reflects the customer's cost of electricity, we conclude that the customer saves about $3\%$ on the electricity costs associated with the thermostat. The lower total energy costs for higher energy consumption was also seen in a related ``smart-device'' demonstration project\cite{Olympic}.
Note, that the quantitative conclusions drawn and numbers presented above were meant to illustrate the questions the MDP approach can resolve. The conclusions and the numbers do not represent any real device as the parameters used were not justified by actual data.
\section{Discussions, Conclusions and Path Forward}
\label{sec:Path}
To conclude, we have presented a novel modeling framework to analyze future demand response technologies. The main novel aspect of our approach lies in the capability of the framework to describe behavior of the smart devices under varying/fluctuating electricity prices. To achieve this goal, we modeled the devices as rational agents which seek to maximize a predefined reward function associated with its actions. In general, the reward function includes the price paid for the electricity consumption and the level of owner discomfort associated with the choices made by the device. At the mathematical level, the system can be described via Markov Decision Processes that have been extensively studied over the last 50 years. Utilizing the MDP approach, we showed that a great variety of practical devices can be described within the same framework by simply changing the set of device states, actions and reward functions. Specifically, we identified four main device categories and proposed simple MDP models for each of them. These four categories include optional loads (like light dimming), deferrable loads (like dishwashing), control loads (thermostats and ventilation systems), and finally storage loads (charging of batteries).
To illustrate the approach we experimented with a simple model of a smart heating thermostat. The MDP-optimized policy of the thermostat followed the expected pattern: it chooses to not heat or keep the temperature stationary at high prices and prefers to heat when the price is low. This policy resulted in $3\%$ of savings in the price paid for electricity, but at the same time led to the total of $3\%$ increase in the consumption level due to the energy costs associated with the thermostat actions. The resulting demand curve showed a noticeable amount of elasticity, thus meeting the main objective of the demand response technology.
There are many relevant aspects of the model that we did not discuss in the manuscript. We briefly list some of these and future research challenges and direction.
\begin{itemize}
\item \emph{Learning algorithms}. In our model we assumed that smart devices have an accurate model of stochastic dynamics for external factors (such as price for electricity), and use this model to find the optimal policy. In reality, however, this model is not known ab initio and has to be learned from the observations. Moreover, one can expect that the dynamics of external factors will be highly non-stationary (i.e. the transition matrix $T(s^{(e)}_{t+1}|s^{(e)}_{t+1})$ will have an explicit dependence on time). Therefore, the optimal policy has to be constantly adapted to the varying dynamics of the external factors. Of a special practical interest is the generalization of the framework to almost periodic processes, reflecting natural daily/weekly/yearly cycles in the electricity consumption.
\item \emph{Price-setting policies}. We did not discuss the price setting policies above, assuming that the policies are given/pre-defined. However, the electricity providers might adjust their policies to consumer response. As the electricity providers pursue their own goals, this setting essentially becomes game-theoretic and as such it requires more sophisticated approaches for analysis. Another extension of the model is to introduce auction-based price-setting schemes, such as in the Olympic Peninsula project \cite{Olympic}. This setting can be naturally incorporated in the same framework, although the modification may require simultaneous modeling of multiple (ensemble of) devices.
\item \emph{Time delays}. Another aspect of the real world not incorporated in our analysis concerns the separation of the time scales associated with operations of the device and intervals of the price variations. Multiple time-scale can be naturally incorporated in the framework by introducing additional states of the device. These modifications will certainly affect final answer for the optimal policy, and the resulting demand curve. However, accurate characterization of the multi-scale behavior will be a challenging task, requiring analysis of nonlinear response functions and dynamical description of the underlying non-Markovian processes.
\end{itemize}
\section{Acknowledgements}
We are thankful to the participants of the ``Optimization and Control for Smart Grids" LDRD DR project at Los Alamos and Smart Grid Seminar Series at CNLS/LANL for multiple fruitful discussions.
\bibliographystyle{IEEETran}
|
\section{Introduction}
Perturbative Quantum Chromodynamics (pQCD) at high energies generates scale called saturation scale $Q_s$.
The need for the existence of such a scale originates back to investigations of unitarity violation \cite{Gribov:1984tu} by linear equations
of pQCD \cite{Kuraev:1977fs,Balitsky:1978ic}.
To resolve this problem, higher order perturbative corrections of nonlinear type within the Balitsky-Kovchegov, CGC/JIMWLK framework (i.e. nonlinear modifications to summation of logarithms of energy $\alpha_s^n(\ln s)^n$) were considered \cite{Balitsky:1995ub,Kovchegov:1999yj,Kovchegov:1999ua,JalilianMarian:1997jx,JalilianMarian:1997gr,JalilianMarian:1997dw,JalilianMarian:1998cb,Kovner:2000pt,Weigert:2000gi,Iancu:2000hn,Ferreiro:2001qy}. These unitarity corrections have a clear physical meaning for they allow gluons to recombine what at some point balances gluons splitting. The saturation scale depends on energy and its existence prevents gluon densities from rapid growth. Existing data suggest that the phenomenon of saturation occurs in nature. The seminal example is provided by a discovery of the geometrical scaling in HERA data \cite{Stasto:2000er} and more recently by a presence of geometrical scaling in production of inclusive jets in the LHC data \cite{McLerran:2010ex,Praszalowicz:2011tc}.\\
Here, we would like to use the saturation property of gluons to derive an expression for amount of thermodynamical entropy associated with generation of saturation scale $Q_s$ which acts as effective mass of gluon during collision of hadrons. We show that such a construction is possible due to the relation of temperature to the saturation scale and the property of saturation scale which acts as an effective mass of gluon. Our main result are formulas (\ref{eq:entropia}), (\ref{eq:finalfinal}), for other approaches to entropy production we refer the reader to \cite{Blaizot:2005wr,Bialas:2006mg,Blaizot:2006tk,Lublinsky:2007mm,Gubser:2008pc}.
We also show that the number of gluons whose distribution is given by unintegrated gluon density (which in perturbative framework is obtained by summing via integral equations Feynman diagrams describing emissions of gluons) (\ref{eq:numberofgluons}) is up to a constant factor equal to number of produced gluons therefore it can be directly linked to entropy of produced gluons. This is possible because of universality of saturation scale which is the only scale in the problem and which links those quantities in a unique way. \\
The letter is organised as follows. In section 2 we obtain a formula for a gluon density in the adjoint representation of the color group and we show that saturation is reached at larger scales than in the fundamental representation. In section 3 we use the obtained gluon density to calculate the cross section for inclusive gluon production and we show that the cross section behaves like the saturation scale. In the 4'th section we derive formula for entropy associated with produced gluons and we relate it to number of gluons given by unintegrated gluon density. Finally we conclude that since saturation scale saturates itself the entropy of gluons coming from saturated region has to be bounded.
\section{Saturation scale in the adjoint color representation}
The high energy factorisable \cite{Catani:1990eg} unintegrated gluon density can be linked to a dipole amplitude for a color dipole scattering off a hadronic target via the following relation \cite{Kovchegov:2001sc,JalilianMarian:2005jf,Kharzeev:2003wz}:
\begin{equation}
\phi_{(G),(Q)} (x,k) \, = \, \frac{C_F}{\alpha_s \, (2 \pi)^3} \, \int d^2 \mathbf b \,
d^2 \mathbf r \, e^{- i \mathbf k \cdot \mathbf r} \ \nabla^2_r \, N_{(G),(Q)} (r,
b, x).
\label{eq:transf}
\end{equation}
where the subscripts $Q, G$ specify whether considered dipole is in the fundamental or in the adjoint color representation. The impact parameter at which a dipole collides with the hadron is $\mathbf b$, the size of the dipole is $r\!=\!|\mathbf r|$ and $k\!=\!|\mathbf k|$ is its conjugated variable characterising transversal momentum of the gluon. The two dimensional vectors $b\!=\!|\mathbf b|$, $k\!=\!|\mathbf k|$, $r\!=\!|\mathbf r|$ lay in the transversal plane to the collision axis. The variable $x$ is related to the longitudinal momentum component of the gluon momentum.\\
Due to the fact that in an electron-proton (e-p) collisions a dipole originates from a photon dissociation the associated gluon density is taken to be in the color fundamental representation while in hadron-hadron (A-A, p-A) collisions it originates most likely from gluon, therefore the associated gluon density with a gluon dipole is in the color adjoint representation. The relation between these two amplitudes in the large $N_c$ limit is the following \cite{Kovchegov:2001sc} :
\begin{equation}
N_G ({r}, { b}, y) \, = \, 2 \, N_Q ({ r}, {b}, y) - N_Q^2
({ r}, {b}, y),
\label{eq:2NN}
\end{equation}
In our investigations of the gluon density in the adjoint representation, we use the one which can be obtained from the GBW model \cite{GolecBiernat:1998js} for the dipole amplitude. This model although, quite simplistic, is a nice laboratory for studying physics of saturation. In particular, we find that the integrals needed in investigated in the next section inclusive gluon production can be performed exactly.
The GBW amplitude reads:
\begin{equation}
N_Q(x,r,b)=\theta(b_0-b)\Bigg[1-\exp\left(-\frac{r^2 Q_s^2(x)}{4}\right)\Bigg]
\label{eq:GBW98}
\end{equation}
where $Q_s(x)$ is the saturation scale and it is modelled to be $Q_s(x)\!=\!Q_0\left(\frac{x_0}{x}\right)^{\lambda/2}$ where $\lambda$, $x_0$, are free parameters and $b_0$ defines proton's radius. This amplitude saturates for large dipoles $r\!>\!\!>\!\!2/Q_s$ and exhibits geometrical scaling which has been confirmed by data \cite{GolecBiernat:1998js,Stasto:2000er}.
The unintegrated gluon density in the adjoint representation can be obtained in a straightforward way via the transformation (\ref{eq:transf}).
However, before we perform this transformation, let us recall the result for the gluon density in the fundamental representation:
\begin{equation}
\phi_Q (x,k)=\frac{N_c A_{\perp}}{2\pi^2\alpha_s}\frac{k^2}{Q_s(x)}\exp\left[-\frac{k^2}{Q_s^2(x)}\right],
\label{eq:ugd2}
\end{equation}
where we assumed a cylindrical shape of the proton which integrated over the impact parameter $b$ gives the transversal area $A_{\perp}=\pi R^2$, where $R$ is a radius of a proton.
Now motivated by the fact that this function exhibits a maximum as a function of $k^2$, see fig. (\ref{fig:plotvel})
for fixed $x$, we define the saturation scale as momentum for which the $\phi_Q(x,k)$ has maximum as a function of $k^2$:
\begin{equation}
Q_s(x)\equiv\partial_{k^2} \phi_Q(x,k)=0
\label{eq:poch2}
\end{equation}
Applying this condition we obtain:
\begin{equation}
\partial_{k^2}\phi_Q(x,k)=\frac{\phi_Q(x,k_t)(Q_s^2(x)-k_t^2)}{k^2Q_s^2(x)}=0
\label{eq:poch2}
\end{equation}
This observation has been used in \cite{Kutak:2009zk} to formulate the GBW model as a solution to a transport equation of a form:
\begin{equation}
\partial_Y \phi(Y,L)+\lambda\partial_L \phi(Y,L,b)=0
\label{eq:trans}
\end{equation}
with initial condition $\phi(k^2,x_0)=k^2 \exp(-k^2)$
where $Y=\ln 1/x$ and $L=\ln k^2$.
Now we calculate the saturation scale for gluons in the color adjoint representation. To see that the effective saturation scale for gluons in the adjoint representation is larger than in the fundamental representation we write the generic expression for the dipole cross section which saturates at large dipole size as:
\begin{equation}
N_Q(x,b,r)=1-e^{-f(x,r,b)}
\label{eq_gludipol}
\end{equation}
Using (\ref{eq:GBW98}) in (\ref{eq:2NN}) and rearranging terms we obtain:
\begin{equation}
N_G(x,b,r)=1-e^{-2f(x,r,b)}
\end{equation}
where the factor $2$ in the exponent is the origin of the difference between the saturation scale for gluon density in the adjoint representation and gluon density in the fundamental representation.
To obtain the gluon density we use (\ref{eq:transf}) together with (\ref{eq:2NN}) to obtain gluon density which reads:
\begin{equation}
\phi_G(x,k)=\frac{C_FA_{\perp}}{4\pi^2\alpha_s}\frac{k^2}{Q_s^2}\exp\left[-\frac{k^2}{2Q_s^2(x)}\right]
\label{eq:adjointglue}
\end{equation}
Applying our definition for saturation scale and calculating derivative we find:
\begin{equation}
\partial_{k^2}\phi_Q(x,k^2)=\frac{\phi_G(x,k)(2Q_s^2(x)-k^2)}{2k^2 Q_s^2(x)}
\label{eq:poch2}
\end{equation}
what means that the effective saturation scale for gluons in the adjoint representation is larger than in the fundamental representation
and reads $(Q_s^2)_G=2(Q_s^2)_Q$.
\begin{figure}[t!]
\begin{picture}(30,30)
\put(30, -80){
\special{
psfile=inclusive.eps
hscale=90
vscale=90
angle=0}
}
\end{picture}
\vspace{3cm}
\caption{\em \small High energy factorisable gluon density. The maxima signalise existence of the saturation scales. Magenta continuous line represents the gluon density in the adjoint representation. The blue dashed line represents the gluon density coming from the dipole in the fundamental representation. Plot is made for $x=10^{-4}$.}
\label{fig:plotvel}
\end{figure}
\section{Inclusive gluon production}
In this section we calculate the rapidity dependence of the cross section for the inclusive gluon production using the gluon density (\ref{eq:adjointglue}). Since we will use only the gluon density in the adjoint representation we skip the index $G$. Here, we consider a slightly asymmetric configuration near the central region, that is why we use the same parameterisation for the gluon densities but with two different saturation scales. In the following, we are interested in qualitative results, therefore for simplicity, we assume hadrons to be protons. Following \cite{Kharzeev:2004if}, we use $x_1=(Q_{s\,1}/\sqrt{s})e^{y}$ and $x_2=(Q_{s\,2}/\sqrt{s})e^{-y}$ where $\sqrt{s}$ is the energy of the collision.
The formula for inclusive gluon production reads \cite{Gribov:1984tu,Kovchegov:2001sc,Braun:2000bh,Kharzeev:2003wz,Kovner:2006wr}:
\begin{equation}
\frac{d \sigma}{d^2 p_t \, dy} \, = \, \frac{2 \, \alpha_s}{C_F} \,
\frac{1}{{p_t}^2} \, \int d^2 k \, \phi({ k}, x_1) \, \phi
({{k} - p_t}, x_2)
\label{eq:inclusiveg}
\end{equation}
For recent theoretical studies of the inclusive gluon production with the NLO effects coming from the running coupling constant we refer the reader to \cite{Horowitz:2010yg}. For recent phenomenological works see \cite{Levin:2010br,Levin:2010zy,ALbacete:2010ad,Albacete:2010bs,Tribedy:2011yn,Tribedy:2010ab,Dumitru:2010iy,Dusling:2009ni}.
Using the expressions for the unintegrated gluon densities and performing the integral over $d^2k$, we obtain:
$$
\frac{d\sigma}{dy\,d^2p_t}=\frac{A_{\perp}^2C_F Q_{s\,1}^2 Q_{s\,2}^2 e^{-\frac{p_t^2}{Q_{s\,1}^2+Q_{s\,2}^2}}}{2 \pi ^3\alpha_s
(Q_{s\,1}^2+Q_{s\,2}^2)^5}
(p_t^4 Q_{s\,1}^2 Q_{s\,2}^2+p_t^2 (Q_{s\,1}^2-Q_{s\,2}^2)^2 (Q_{s\,1}^2+Q_{s\,2}^2
$$
\begin{equation}
+2 Q_{s\,1}^2 Q_{s\,2}^2
(Q_{s\,1}^2+Q_{s\,2}^2)^2))
\label{eq:rapdependence}
\end{equation}
Integrating over $d^2p_t$ and introducing minimal $p_{t\,min}$ because the integral is logarithmically divergent, we obtain:
\begin{equation}
\frac{d\sigma}{dy}=\frac{2A_{\perp}^2 C_F Q_{s\,1}^2 Q_{s\,2}^2 e^{-\frac{p_{t\,min}^2}{Q_{s\,1}^2+Q_{s\,2}^2}} \left(p_{t\,min}^2 Q_{s\,1}^2 Q_{s\,2}^2+Q_{s\,1}^6+Q_{s\,2}^6\right)}{\pi ^2 \alpha_s (Q_{s\,1}^2+Q_{s\,2}^2)^4}
+
\frac{4A_{\perp}^2 C_F Q_{s\,1}^4 Q_{s\,2}^4 \Gamma \left(0,\frac{p_{t\,min}^2}{(Q_{s\,1}^2+Q_{s\,2}^2)}\right)}{\pi ^2 \alpha_s
(Q_{s\,1}^2+Q_{s\,2}^2)^3}
\label{eq:crossrapid}
\end{equation}
Assuming, that the saturation scale in one of the colliding hadrons is the smallest scale in the problem $p_{t\,min}\!\simeq\!Q_{s\,1}\!\!<\!\!Q_{s\,2}$ we perform expansion in terms of $Q_{s\,1}/Q_{s\,2}$ in (\ref{eq:crossrapid}) arriving at:
\begin{equation}
\frac{d\sigma}{dy}=\frac{2 A_{\perp}^2 C_F Q_{s\,1}^2}{\pi ^2 \alpha_s}-
\frac{2 A_{\perp}^2 C_F Q_{s\,1}^4}{\pi ^2 \alpha_s
Q_{s\,2}^2}\left(2 \log \left(\frac{Q_{s\,1}^2}{Q_{s\,2}^2}\right)+5+2 \gamma_E\right)+O\left((Q_{s\,1}^2/Q_{s\,2}^2)^3\right)
\label{eq:finaleq}
\end{equation}
$$
$$
A comment regarding formula (12) is in order. In derivation of (12) it has been assumed that one of gluon densities is in a dilute regime, the more general formulation which would allow for taking into account interaction of two dense targets would most probably introduce regulator of the infrared divergence. However, such framework at present is not known and we regulate the divergence by the lowest physical scale in the problem.
The result above shows in agreement with \cite{Horowitz:2010yg,Levin:2010br} that $d\sigma/dy$ behaves like the saturation scale making this observable particularly interesting for studying physics of dense partonic systems formation. On Fig. (\ref{fig:plotvel}) we show results obtained from formulas
(\ref{eq:crossrapid}) and (\ref{eq:finaleq}). We see that the full result is well approximated by the first term if the saturation scales are significantly different.
\begin{figure}[t!]
\begin{picture}(30,30)
\put(30, -80){
\special{
psfile=inclusivegluonproduction_plots.eps
hscale=90
vscale=90
angle=0}
}
\end{picture}
\vspace{3cm}
\caption{\em \small Cross section for inclusive gluon production (normalised to larger of two saturation scales and square of the transversal area $A_\perp\approx S_\perp$) as a function of ratio of saturation scale. The blue line is the full cross section (\ref{eq:crossrapid}), the magenta line is the leading term, the green is the contribution from the sum of leading and subleasing term.}
\label{fig:plotted}
\end{figure}
\section{Entropy}
Now we would like to relate the obtained cross section for the inclusive gluon production to produced entropy in the collision due to emergence of effective mass of gluons during the collision.
We wish, however, to consider dense system where the following formula is not easily justified :
\begin{equation}
S=\int\frac{d^3z\,d^3k}{(2\pi)^3}[-f(z,k)\ln f(z,k)+s_I(1+s_if(z,k))\ln(1+s_if(z,k))]
\end{equation}
where $f(z,p)$ is phase space density of patrons/hadrons, $s_i=-1$ for fermions and $s_i=1$ for bosons.
We see the possibility to address the question of entropy of dense system, following ideas of \cite{Kharzeev:2005iz,Castorina:2007eb} where relation between saturation scale and temperature of thermally produced hadrons has been obtained within partonic picture of QCD and it reads:
\begin{equation}
T=\frac{Q_s(x)}{2\pi}
\label{eq:temperatura}
\end{equation}
where $x$ stands for a longitudinal momentum fraction of hadron carried by a parton.
The formula above is a direct consequence of the Unruh effect \cite{Unruh:1976db} within QCD. The Unruh effect \cite{Unruh:1976db} states that accelerated observer in its rest frame feels thermal radiation of temperature $T=\frac{a}{2\pi}$. In the construction leading to QCD application, a decelerating off shell gluon with distribution in transversal momentum given by unintegrated gluon distribution is considered. Due to deceleration, a horizon is formed tunnelling through which produces a thermal ensemble with a density matrix of the Maxwell-Boltzmann type. To determine the deceleration the authors of \cite{Kharzeev:2005iz} use the Wong equations \cite{Wong:1970fu} for color particle moving in a homogeneous chromoelectric field. Solutions of this equations give a result that $|a|=Q_s$ which leads to
eq. (\ref{eq:temperatura}).
Furthermore, it is well known that saturation scale effectively acts as a regulator of the infrared divergent behaviour of the unintegrated gluon density. At the lowest order in QCD (in phase space region where the total energy is larger than any other scale) it is:
\begin{equation}
\phi=\frac{\alpha_s C_F}{\pi}\frac{1}{k^2}
\end{equation}
This quantity has a structure of propagator and we see that at small values of $k$ it becomes divergent.
Higher order perturbative corrections coming from BFKL resumation do not solve this problem. This low $k$ behaviour of BFKL equation is interpreted as leading to infinitely many gluons with small $k$ and also to power like growth of parton density with energy what violates unitarity bounds. Saturation due to nonlinear effects modifies the behaviour of distribution of gluons and in particular, since the unintegrated gluon density has a maximum, it sets the most probable momentum to be of the order of $Q_s$, what can be read off from formula (\ref{eq:poch2}). This we interpret as giving by nonlinear effects (here we study saturation model but saturation is also studied within nonlinear QCD evolution equations \cite{Kovchegov:1999yj,Kovchegov:1999ua}) emergent mass (gauge invariance still holds) to gluon which is proportional to the saturation scale and which renders the propagator to be finite.\\
We are also motivated by studies of QCD by means of nonperturabative methods where it is argued that gluon mass consistent with gauge invariance might be dynamically generated due to nonlinearities (see \cite{Mathieu:2011mq} and references therein).
The regulated divergency of propagators due to generation of effective mass $Q_s(x)$ of virtual gluons has also impact on real gluons since they acquire mass. We can therefore write for single real gluon :
\begin{equation}
M_G(x)=Q_s(x)
\end{equation}
This observations allows us to define entropy of gluons, because the existence of temperature and existence of effective mass of gluons allows for introduction of the thermodynamical entropy (coarse grained entropy).
Using the laws of thermodynamics and assuming small change of the volume associated with system of gluons we obtain:
\begin{equation}
dE=TdS
\label{ref:termo}
\end{equation}
and setting $dE=dM$
gives:
\begin{equation}
dM=TdS
\label{eq:termo}
\end{equation}
where the mass is understood to be the emergent mass of system of gluons carrying small longitudinal momentum fraction of the hadron and is related to mass of single gluon via:
\begin{equation}
M(x)=N_G(x) M_G(x)
\end{equation}
where $N_G(x)$ is (dimensionless) number of gluons defined as:
\begin{equation}
N_G(x)\equiv\frac{dN}{dy}=\frac{1}{S_\perp}\frac{d\sigma}{dy}
\end{equation}
where $dN/dy$ is multiplicity of produced gluons.
Now we use eq (\ref{eq:temperatura}) what allows us to link the saturation scale to produced entropy:
\begin{equation}
d\left[N_G(x)\,M_G(x)\right]=\frac{Q_s(x)}{2\pi}dS
\label{eq:entropianowa}
\end{equation}
Thus we obtain:
\begin{equation}
S=\frac{6C_F \,A_{\perp}}{\pi\alpha_s}Q_s^2(x)+S_0
\label{eq:entropia}
\end{equation}
and equivalently
\begin{equation}
S=3\pi\,\left[N_G(x)+\,N_{G\,0}\right]
\label{eq:finalfinal}
\end{equation}
where $S_0=\frac{6C_F\,A_{\perp}}{\pi\alpha_s}\mu^2$ is a state of the lowest entropy or the minimal number of gluons.\\
Obtained entropy can be also expressed in terms of number of gluons whose distribution is given by unintegrated gluon density. The number of gluons which build up unintegrated gluon density is:
\begin{equation}
n_G(x)\equiv\frac{1}{\pi}\int d^3r\,d^2k\,\Phi(x,k,r)=\frac{1}{\pi}\int d^2k\,\phi(x,k^2)
\label{eq:numberofgluons}
\end{equation}
where $d^3r\equiv d^2b\,dl$ and $r\equiv(l,{\bf b})$, $l$ is longitudinal dimension of the proton, $b$ is already introduced impact parameter, $\Phi(x,k,r)$ is understood to
be gluon density unintegrated in $r$, $k$, $x$.
For simplicity we assume that longitudinal space distribution of gluons factorises and is given by $\delta(l)$. This assumption is justified since the nontrivial dynamics takes place in the transversal space. Therefore we obtain:
\begin{equation}
n_G(x)=\frac{C_F A_\perp}{2\pi^2\alpha_s}Q_s^2(x)
\end{equation}
Now we can expressed entropy of produced gluons by number of gluons given by unintegrated gluon density:
\begin{equation}
S=12\pi\,n_G(x)+3\pi\,N_{G\,0}
\label{eq:final2}
\end{equation}
This can be used perhaps to define via (\ref{eq:final2}) entropy of gluons described by unintegrated gluon density, however we postpone this task for future studies.
Although, we used GBW model to arrive at formula (\ref{eq:entropia}) our result is model independent in saturated region since GBW model is consistent with results of\cite{McLerran:1993ni,Kharzeev:2003wz,Kovchegov:1999yj,Kovchegov:1999ua,Iancu:2003ge} where gluon density in saturated region behaves like:
\begin{equation}
\phi(x,k^2)\sim \frac{C_F A_\perp}{\alpha_s}\frac{k^2}{Q_s^2(x)}
\label{eq:mclarren}
\end{equation}
The results we obtained above and in particular (\ref{eq:entropia}) are valid when $Q_s\!\!>\!\!>\mu$ where $\mu$ is an unknown constant, we choose this constant to be $\Lambda_{QCD}$. In order to obtain a formulation valid also for small temperatures one has to go along lines of the Nernst theorem,
going along these lines, one can expect that the lowest possible entropy will be zero what will correspond to situation where coherence of a proton \cite{Good:1960ba} is not affected by interactions.
Furthermore, we see in agreement with intuition that the entropy is proportional to number of gluons and depends linearly on the hadron size i.e. doubling the hadron size the entropy is doubled.\\
Here we would like to stress that our construction of entropy given by (\ref{eq:finalfinal}) is quite general. The crucial element is the existence of saturation which links the temperature and the mass of system of gluons allowing for introduction of thermodynamical relations between these quantities.\\ There is one more consequence of the dependence of entropy on the saturation scale. It gives upper bound on the amount of produced entropy associated with gluons confined in saturated region of phase space what follows from our result combined with result of \cite{Albacete:2008ze,Avsar:2010ia}. In these investigations authors found that at extreme energies saturation scale eventually saturates what in combination with (\ref{eq:entropia}) gives:
\begin{equation}
\Delta S_{max}=3\pi\,\left[N_{G\,max}(x)+\,N_{G\,0}\right]
\label{eq:entropia2}
\end{equation}
Entropy however can still be produced outside of saturated region even if saturation scale will saturate itself. This we can guess from approaches which take care for more realistic behaviour of gluon density outside of saturation region.
Namely outside of the saturation region i.e. $k>Q_s$ the gluon density from GBW model falls too fast (due to exponential behaviour it can be integrated to infinity) while BK or McLerran-Venugopalan \cite{McLerran:1993ni} model gives hard perturbative behaviour of type $Q_s^2/k^2$ which gives logarithmic divergent contribution after integrating what will give contribution to entropy unbounded by saturation of the saturation scale.
\section{Conclusions}
In this letter we have studied properties of the gluon distribution possessing the feature of saturation. We defined the saturation scale mathematically as a maximum of the gluon density (\ref{eq:poch2}). We have calculated the cross section for the inclusive gluon production and have shown that it might be approximated by saturation scale characterising gluon density of less saturated hadron. Furthermore, we introduced saturation scale related entropy of produced gluons and expressed it in terms of multiplicity of produced gluons. We have shown that this entropy behaves like number of gluons whose distribution is given by unintegrated gluon density.
We also concluded that since saturation scale ultimately saturates itself, there should exist an upper bound on entropy of gluons coming from saturated part of gluon distribution. The result we obtained is of course a contribution to the entropy coming just from emergence of saturation scale and effective mass of gluons in perturbative regime.\\
\section*{Acknowledgements}
I would like to thank Krzysztof Golec-Biernat and Andrzej Bialas for comments and interesting discussions. Also discussions with Wojciech Florkowski, Piotr Surowka, Eddi de Wolf are kindly acknowledged. Finally I would like to thank Yuri Kovchegov for useful correspondence.\\
This work has been supported by {\bf Fundacja na rzecz Nauki Polskiej} with grant {\bf HOMING PLUS/2010-2/6}.
|
\section{Introduction}
\label{sec:introduction}
The matter power spectrum is a sensitive probe of cosmological models.
Observations of galaxy clustering, weak gravitational lensing, and the Lyman-$\alpha$ forest are all tracers of the matter power spectrum and have to-date provided key constraints on the initial conditions of the universe as well as on the growth and expansion history. Many future surveys will rely on tracers of the matter power spectrum to constrain models for dark energy, dark matter, and massive neutrinos. However, due to nonlinear gravitational evolution of the matter distribution, both the mean matter power spectrum as well as its variance and correlations between bands in wavenumber are difficult to predict without expensive numerical simulations.
Considerable work has been done to predict the mean nonlinear matter power spectra to the precision necessary for near-term surveys~\citep[e.g.,][]{smith03, lawrence10}. In order to use observations of the matter power spectrum to constrain cosmological parameters it is necessary to know the error distribution for the power spectrum estimates in addition to the mean spectrum predicted by the cosmological model. It is by now well established that nonlinear gravitational evolution creates a complicated error distribution for the power spectrum with signficantly increased small-scale variance and correlations between all band powers when compared with the Gaussian case~\citep{scoccimarro99, meiksin99,hamilton06,neyrinck06,neyrinck07,takahashi09}. The differences between the Gaussian error and nonlinear error models for the matter power spectrum are quite significant for parameter estimation from galaxy surveys~\citep{hamilton06, neyrinck07}, but uncertainty in the nonlinear galaxy bias with respect to the dark matter may dominate the inferred parameter constraints in the near future. Weak lensing surveys probe the matter distribution directly and are thus insensitive to biasing. The line-of-sight projection in weak lensing reduces the nonlinear contribution to the power spectrum covariance, but it is still a significant consideration for parameter estimation~\citep{semboloni07, takada09, sato09}.
Standard estimators of the matter power spectrum covariance matrix
in a given model require a large ensemble of
statistically independent realizations of the matter density over volumes at least as large as the
survey being analyzed. \citet{meiksin99} found that at least several hundred realizations are
necessary to estimate the covariance over an interesting dynamical range and~\citet{takahashi09}
have used 5000 simulations to obtain a low-noise estimator for a Gpc$^{3}$ volume.
\citet{hamilton06} attempted to estimate the covariance matrix by sub-sampling a single
simulation volume, but found that the window functions they applied to define different
pseudo-independent subvolumes significantly altered the covariance. That is, the Fourier transform
of a windowed density field has Fourier modes that are linear combinations of the modes in the
$N$-body simulation volume with periodic boundary conditions. The modes of the windowed density
then have different covariance than the modes in the periodic box, which is particularly enhanced
by the correlations between the largest and smallest scales~\citep{rimes06}.
In a similar line of inquiry, \citet{norberg08} showed that jacknife estimates of the covariance
matrix from simulated survey regions systematically underpredict the true nonlinear covariance
matrix. Finally, if some information is known about the structure of the power spectrum covariance,
then the ``shrinkage estimator'' of~\citet{pope08} gives a way to reduce both the bias and noise
in covariance estimates from a fixed number of density realizations. However, in our experience the
``shrinkage estimators'' as formulated in \citet{pope08} rely quite heavily on the prior information
one is able to supply.
In this paper we describe a method to obtain multiple realizations of the matter density from a
single $N$-body simulation by resampling those Fourier modes (in a periodic box) that can be
approximated as Gaussian distributed with zero correlations between Fourier modes. While
we anticipate this algorithm could be useful for a number of applications~\citep[such as adding
power larger than the simulation volume as in][]{cole97}, in this paper we focus on the
minimum number of simulations needed to estimate the matter power spectrum covariance matrix.
As a benchmark for our study we use the results of~\citet{takahashi09}, who used 5000 $N$-body
simulations to estimate the covariance matrix in a 1~($h^{-1}$Gpc)$^{3}$ simulation volume.
While we focus here on the 3-D matter power spectrum, one final goal will be to efficiently
estimate the weak lensing power spectrum covariance for upcoming cosmic shear surveys.
An example of this using a brute-force approach is given in~\citet{sato09}.
The structure of this paper is as follows. In Section~\ref{sec:resampling} we describe our method
for resampling large-scale modes in a periodic $N$-body simulation and the necessary adjustments
to the remaining small-scale modes.
We give an approximate algorithm requiring only a single snapshot, but limited to simulations
of at least several Gpc on a side, in Section~\ref{sec:singlesnapshotalgorithm}.
We describe our full algorithm for arbitrary simulation sizes (as long as the large-scale modes
are Gaussian distributed) in Section~\ref{sec:fullalgorithm}.
In Section~\ref{sec:validation} we present validation
tests for adding new large-scale mode realizations.
We describe how our resampling algorithm can be useful for
power spectrum covariance matrix estimation in Section~\ref{sec:covestimation}.
We discuss several future directions and
applications for this work in Section~\ref{sec:conclusions}.
Where numerical results are given, unless otherwise stated,
we assume the same cosmology as~\citet{takahashi09} with
$\Omega_{m}=0.238$, $\Omega_{b}=0.041$, $\Omega_{\Lambda}=0.762$,
$n_{s}=0.958$, $\sigma_{8}=0.76$, and $H_{0} = 73.2~{\rm km}\,{\rm s}^{-1}\,{\rm Mpc}^{-1}$.
\section{Resampling large-scale power}
\label{sec:resampling}
Our goal is to isolate and resample those Fourier modes in a single periodic $N$-body simulation
that can be accurately approximated as Gaussian distributed.
We assume that the Fourier modes of the initial conditions
for the simulation are generated from a random realization of a complex Gaussian field on a grid from a
specified power spectrum ({\it i.e.} the real and imaginary parts of the Fourier modes are each drawn from
independent Gaussian distributions such that the amplitude of the modes is equal to the square root of the
power spectrum). As the simulation is evolved, gravitational interactions between the particles tracing the
matter density field cause the phases of the Fourier modes to become correlated and the distribution
of Fourier mode amplitudes to become skewed towards larger values. The growth rate of correlations
and departure from Gaussianity are functions of wavenumber as gravitational collapse proceeds
most rapidly on small-scales.
So, we first need to separate the Fourier modes into ``large-scales'' defined to be Gaussian distributed, and
``small-scales'' that we will define to be all the remaining Fourier modes in the simulation. Because it
worked for our purposes we crudely separate large and small scale modes by defining
all modes with wavenumber amplitude less than some value $k_{\rm thresh}$ as Gaussian distributed.
In practice we defined $k_{\rm thresh}$ so that the r.m.s. mass overdensity smoothed in top-hat spheres
of radius $2\pi/k_{\rm thresh}$ was less than 0.2~\citep[similar to][]{cole97}. This is based on the idea that larger
mass overdensity fluctuations indicate more nonlinear growth leading to larger departures from
Gaussianity.
Having defined those Fourier modes that we will model as Gaussian, we next resample them
by generating new realizations of the initial conditions
that have nonzero Fourier modes only for $|\kv| < k_{\rm thresh}$
and extrapolating the amplitudes of the modes to the desired redshift using linear theory.
To obtain a new realization of the density field that approximates what
we would have obtained had we run a simulation with the new large-scale modes in the
initial conditions of the simulation,
we have to adjust the small-scale modes in some way to account for the
phase correlations that would have emerged between the large and small scales.
\subsection{Resampling algorithm}
\label{sec:modeaddition}
In this Section we imagine that we know the particle positions in a simulation
that has no large-scale power (where ``large-scale''
denotes Fourier modes with $|\kv| \leq k_{\rm thresh}$) and we seek an algorithm to add
new large-scale modes. We will denote the simulation without large-scale power
as a ``small-modes'' simulation and the simulation with all Fourier modes nonzero as
an ``all-modes'' simulation.
\subsubsection{Physical picture}
Altering the amplitudes and phases of the large-scale modes of the mass density in a
cosmological volume has two dominant effects,
\begin{enumerate}
\item sub-volumes are expanded or compressed,
\item sub-volumes gravitationally evolve with an effective
matter background density that is enhanced or diminished,
\end{enumerate}
depending on where the large-scale modes become more or less overdense,
respectively~\citep{frenk88,little91,tormen96, cole97,angulo10}.
We follow~\citet{cole97} to model both of these effects.
To model the expansion and compression of
sub-volumes, we perturb the particle positions using the
Zel'dovich displacements~\citep{zeldovich70}
calculated from the large-scale modes, $\delta_L$,
\begin{equation}\label{eq:zeldovich}
\dvec_{L}(\xv,a) \equiv \int \frac{d^3k}{(2\pi)^3}
\frac{\delta_L(\kv,a)\kv}{k^2}\,
e^{i\kv\cdot\xv}.
\end{equation}
The expansion and compression of sub-volumes is analogous to the
frequency modulation of a radio wave. The large-scale modes to be
added are a new signal that modulates the sizes of sub-volumes in
the matter density field; expressed to first order by the
Zel'dovich displacements in equation~(\ref{eq:zeldovich}).
Compressed (expanded) regions also evolve with an increased (decreased)
effective matter density, altering the growth rate of structure.
Because the growth rate increases monotonically with time, we
model the changing effective local matter density
by finding new times $a'(\xv)$ at which the linear growth rate matches
the growth rate in a universe with matter density
$\Omega_{m}\left(1+\delta_{L}(\xv, a)\right)$ (with $\Omega_{\Lambda}$ kept fixed).
The perturbed scale factors $a'(\xv)$ are then defined by the relation,
\begin{equation}\label{eq:growthmatching}
D(a',\Omega_m) \approx D(a, \Omega_m \left(1+\delta_L(\xv,a)\right) ),
\end{equation}
where $D(a,\Omega_{m})$ is the linear growth function.
In practice, we solve equation~(\ref{eq:growthmatching}) numerically to get $a'$
as demonstrated in Figure~\ref{fg:scalefactor_vs_omegam}.
\begin{figure}
\centerline{
\includegraphics[scale=0.48]{growth_tables_takahashi.pdf}
}
\caption{\label{fg:scalefactor_vs_omegam}Top left: Linear growth function normalized
to one at the time of the simulation initial conditions ($a_{\rm IC}=1/21$) versus scale factor.
Top right: Linear growth function versus matter density, also normalized to unity at the
time of the initial conditions for all values of $\Omega_{m}$. Bottom: scale factor as a
function of matter density at matched linear growth. The vertical dashed line shows the
value of $\Omega_{m}=\Omega_{m}^{0}=0.238$ used in our simulations.
The vertical dotted lines show the values
of $\Omega_{m}^{0}(1\pm4.3\sigma(\delta_{L}))$, where $\sigma(\delta_{L})\approx0.17$ is the r.m.s.
mass over-density in top-hat spheres at the scale $R=2\pi/(256 k_{F})$ ($k_F\equiv 2\pi/L_{\rm box}$) for
the density perturbations with nonzero Fourier amplitudes
only for $|\kv| < k_{\rm thresh}$. The horizontal dotted lines
in the bottom panel show the corresponding scale factor range over which our small-modes
simulation must be saved.}
\end{figure}
\subsubsection{Approximate method for a single snapshot}
\label{sec:singlesnapshotalgorithm}
Combining the two effects of expansion/compression and time-perturbations,
the model for adding large-scale fluctuations to the mass density
field is,
\begin{equation}\label{eq:densitymodel}
\delta(\xv, a) = \delta_{L}(\xv,a) + \delta_{S}(\xv', a'),
\end{equation}
where $\delta(\xv,a)\equiv\rho(\xv,a)/\bar{\rho} - 1$ is the mass over-density that would
be measured in the all-modes simulation, $\delta_{S}(\xv,a)$ is the
mass over-density measured in the small-modes simulation, $\delta_{L}$ has zero power for
$|\kv| > k_{\rm thresh}$, $a'$ is defined by equation~(\ref{eq:growthmatching}),
and $\xv' \equiv \xv + \dvec_{L}(\xv,a)$.
Note that for this definition of $\xv'$ the Zel'dovich displacements
are added to the particle positions prior to application of the time-perturbation
(i.e., at the positions at time $a$, not at $a'$).
We will discuss this choice below.
When $|a' - a| \ll 1$ (which requires $\left< \left|\delta_L(\xv)\right|^{2}\right>^{1/2} \ll 1$)
the first order correction to $\delta_{S}$ in $a'$ should be
sufficient for adding $\delta_{L}$ to the density field.
Expanding both sides of equation~(\ref{eq:growthmatching})
to first order in $(a'-a)$ and $\delta_L$,
\begin{align}
D(a') &\approx D(a) + (a'-a)\frac{dD}{da}
\\
D(a,\Omega_m(1+\delta_L(\xv,a))) &\approx D(a,\Omega_m) +
\Omega_m\delta_L(\xv,a)\frac{dD}{d\Omega_m}.\notag
\end{align}
Solving for $a'$ gives,
\begin{equation}\label{eq:timeperttaylor}
\frac{a'}{a} -1\approx \delta_L(\xv,a)
\left(\frac{d\ln D}{d\ln\Omega_{m}} / \frac{d\ln D}{d\ln a}\right),
\end{equation}
which shows that the perturbed scale factor is roughly proportional to the
large-scale density perturbations being added.
Using equation~(\ref{eq:timeperttaylor})
in the Taylor expansion of equation~(\ref{eq:densitymodel}) gives,
\begin{multline}\label{eq:deltataylor}
\delta(\xv,a) \approx \delta_{L}(\xv,a) + \delta_{S}(\xv',a) + \\
\delta_{L}(\xv,a)\, \frac{\partial\delta_{S}(\xv',a)}{\partial\ln a}\,
\left(\frac{d\ln D}{d\ln\Omega} / \frac{d\ln D}{d\ln a}\right).
\end{multline}
This suggests that the addition of large-scale modes can be quickly
computed given an estimate of the nonlinear growth rate of the small-modes simulation.
With a single snapshot, the growth rate can be estimated from the
continuity equation\footnote{Note that for simulation volumes $\gg 1$~Gpc on a side
relativistic corrections to the continuity equation could be important.
These arise from the way that
the large scale potential perturbations modify the geodesic equations.},
\begin{equation}
H(a)\frac{\partial\delta(\xv,a)}{\partial\ln a} =
-\frac{1}{a}\nabla\cdot
\left((1+\delta)\mathbf{v}\right) ,
\end{equation}
and, using equation~(\ref{eq:densitymodel}),
\begin{equation}
\frac{\partial\delta_S(\xv,a)}{\partial\ln a} =
\frac{\partial\delta(\xv,a)}{\partial\ln a}
- \frac{d\ln D}{d\ln a}\delta_L(\xv,a).
\end{equation}
So, given particle positions and velocities at scale factor $a$ for a simulation
that has no large-scale modes, equation~(\ref{eq:deltataylor}) provides a method
to add new large-scale mode realizations when the large-scale density perturbations
are small.
According to equation~(\ref{eq:timeperttaylor}) the range of $a'-a$ is proportional to the
range of $\delta_{L}(\xv, a)$. Because $\delta_{L}(\xv,a)$ is Gaussian distributed
by construction, we can determine the range of $\delta_{L}$ from its variance,
\begin{equation}\label{eq:sigmatophat}
\sigma^{2}\left(\delta_{L}(\xv,a)\right) \equiv
\int_{k_{F}}^{k_{\rm thresh}} d\ln k\, \Delta^{2}(k)\,
W_{\rm top-hat}^{2}(k R),
\end{equation}
where $k_{F}\equiv 2\pi/L_{\rm box}$ is the fundamental frequency of the simulation box,
$\Delta^{2}(k)\equiv k^3 P(k) / (2\pi^2)$ is the theoretical dimensionless matter
power spectrum, and $W_{\rm top-hat}(k R)$ is the Fourier transform of a spherical top-hat
window of radius $R$.
The scale $R$ is roughly given by either the pixel scale of the gridded density field
or the mean inter-particle spacing, whichever is larger. For the simulations we use
below, $L_{\rm box}=1000$~$h^{-1}{\rm Mpc}$, $k_{\rm thresh}=8 k_{F}\approx 0.05h^{-1}$Mpc, and, with 256$^{3}$
particles, $R\sim 2\pi/(256 k_{F})$.
Evaluating equation~(\ref{eq:sigmatophat}) gives $\sigma\approx 0.17$ which is
still less than $0.2$ as we required when choosing $k_{\rm thresh}$. But, as shown in
Figure~\ref{fg:scalefactor_vs_omegam}, the variations in $\delta_{L}(\xv,a)$ greater than
about $2\sigma$ lead to $|a'-a| > 1$ making equation~(\ref{eq:deltataylor})
insufficiently accurate. In our test simulations $\delta_{L}$ actually varies over
$\pm 4.3\sigma \approx \pm0.87$, giving a range of perturbed scale
factor $0.26 \lesssim a' \lesssim 4.4$ (shown by the dotted lines in
Figure~\ref{fg:scalefactor_vs_omegam}).
In Figure~\ref{fg:avslbox} we show the maximum value of $|a'-a|$ as a function
of the simulation box size while keeping $k_{\rm thresh}/k_{F}$ fixed at 8 and, as in the preceding
paragraph, assuming maximal variations in $\delta_{L}(\xv,a)$ of $4.3\sigma$.
\begin{figure}
\centerline{
\includegraphics[scale=0.48]{max_aprime_vs_lbox.pdf}
}
\caption{Maximum variation in perturbed scale factor $a'$ as a function
of the simulation box size for fixed $k_{\rm thresh}/k_{F}=8$ and assuming the maximum
fluctuations in $\delta_{L}(\xv,a)$ are $4.3\sigma$. Top: Maximum absolute difference
between $a'$ and $a=1$ for the growth matching condition for adding large-scale
modes to a simulation (equation~\ref{eq:growthmatching}).
Bottom: Maximum $a'$ for the growth matching condition
for removing large-scale modes from a simulation (equation~\ref{eq:inversegrowthmatching}).
There is no solution for $a'$ left of the vertical dashed line.}
\label{fg:avslbox}
\end{figure}
For $L_{\rm box} \sim 3000$$h^{-1}{\rm Mpc}$, as in the recently completed MXXL
simulation\footnote{The MXXL simulation (Angulo et al. in preparation)
was run by the Virgo Consortium (\url{http://www.virgo.dur.ac.uk})
in the spring of 2011. It is 3~$h^{-1}$Gpc on a side with $3\times 10^{11}$ particles
with mass $6.3\times 10^9\,M_{\odot}$.},
the variation in $a'-a$ is less than 0.2 and the
approximate formula in equation~(\ref{eq:deltataylor}) could be sufficient for adding the
large-scale modes to the small-modes simulation.
To further assess the validity of equation~(\ref{eq:deltataylor}), we show a histogram of the
values of the third term in equation~(\ref{eq:deltataylor}), normalized by $\delta_S(\xv',a)$,
on a 512$^3$ grid in our 1~$h^{-1}$Gpc small-modes simulation in Figure~\ref{fg:nonlineargrowthhist}.
\begin{figure}
\centerline{
\includegraphics[scale=0.48]{deltaSTaylorCorrectionNormalized.pdf}
}
\caption{Histogram of the values of $\frac{(a'-a)}{a}\,\frac{d\ln\delta_S}{d\ln a}$
on a 512$^3$ grid for our $\Lambda$CDM cosmology in a 1~($h^{-1}$Gpc)$^3$ simulation.
The tick marks on the top (blue) axis show the projected bin values if we
scale the bins on the bottom axis by the ratio of the maximum perturbed scale factor,
${\rm max}(a'-a)$,
in a 3~$h^{-1}$Gpc box versus the maximum perturbed scale factor in our
1~$h^{-1}$Gpc box. The vertical dashed lines show the values of $\pm 1$ on this axis,
where the third term in equation~(\ref{eq:deltataylor}) would be small relative to $\delta_S$.}
\label{fg:nonlineargrowthhist}
\end{figure}
The distribution of grid cell values is strongly peaked around zero, but has wide tails
extending to values of several hundred. This again illustrates that equation~(\ref{eq:deltataylor})
is a bad approximation for a (1~$h^{-1}$Gpc)$^3$ volume. However, if we scale the
bin values on the lower horizontal axis by the ratio of the maximum perturbed scale
factor ${\rm max}(a'-a)$ in a (3~$h^{-1}$Gpc)$^3$ simulation to the maximum $(a'-a)$ in
our simulation, we get the axis on the top of Figure~\ref{fg:nonlineargrowthhist}.
For a (3~$h^{-1}$Gpc)$^3$ simulation 92\% of the grid cells of the third term in
equation~(\ref{eq:deltataylor}) have values less than one. So, equation~(\ref{eq:deltataylor}) may
be a good enough approximation for adding large-scale modes to the MXXL or other
simulations with similarly large volumes.
\subsubsection{Full recipe}
\label{sec:fullalgorithm}
For box sizes smaller than several Gpc on a side, we need an algorithm that
accurately implements the growth matching condition of equation~(\ref{eq:growthmatching})
to compute the density $\delta_S(\xv',a')$ in equation~(\ref{eq:densitymodel}).
For computational expediency and for later comparison
with~\citet{takahashi09}, we have limited our test simulations to
$L_{\rm box}=1000$$h^{-1}{\rm Mpc}$.
Therefore, we have not tested the
accuracy of equation~(\ref{eq:deltataylor}).
Instead, because our test simulation
parameters yield $|a'-a|>1$, we apply the model of equation~(\ref{eq:densitymodel})
by adjusting the positions of each particle in the simulation.
If $\xv_{i}(a)$ is the position of the $i$th particle at scale-factor $a$ in
the small-modes simulation, the model from equation~(\ref{eq:densitymodel})
with modes $\delta_{L}$ added to the simulation
can be rewritten for each particle position as,
\begin{equation}
\xv_{i}'\left(a'; \delta_{L}\right) \equiv \left[ \xv_{i}(a) + \dvec_{L}\left(\xv_{i}(a), a^{*}\right)\right]_{a = a'},
\end{equation}
where $a^*$ is the scale factor of the
target time where the full
density field (with new large-scale modes) is desired.
That is, we first apply the same Zel'dovich displacement to the particle positions in all
the snapshots (evaluated at the unperturbed positions in each snapshot).
Then, we interpolate between these modified snapshots to find the particle position
at perturbed time $a_{i}' \equiv a'(\xv_{i} + \dvec(\delta_{L}(\xv_{i}(a)), a^{*}))$.
To find the particle positions at all values of $a'$ we must save simulation
snapshots at times covering the range of $a'$ values and spaced sufficiently
closely in $a$ so that the interpolation of positions is accurate. This requires
running the simulation far into the future ($a > 4$ for our test simulations) and
saving $\sim 10$'s of snapshots, which creates both an extra computational
and also a disk storage burden over the approximation in equation~(\ref{eq:deltataylor}).
The steps of our full algorithm for adding large-scale modes to the small-scale
density are then:
\begin{enumerate}
\item Choose a threshold scale $k_{\rm thresh}$ defining the boundary between large and small scale modes and a target time $a^*$ where the resampled density is desired.
\item Calculate the range of perturbed scale factor $a'$ defined by
equation~(\ref{eq:growthmatching}) about $a^*$ where the simulation snapshots will be required.
\item Run an $N$-body simulation in a periodic volume with nonzero Fourier modes
in the initial conditions only for $|\kv| > k_{\rm thresh}$ and save snapshots over the required
range of $a'$ that are spaced sufficiently close together to allow for accurate
interpolation of particle positions.
\item Calculate a new realization of large-scale modes $\delta_L(a^*)$ on
a grid.
\item Calculate the Zel'dovich displacements defined in equation~(\ref{eq:zeldovich}) from
$\delta_L$.
\item Apply the Zel'dovich displacements to each snapshot of the small-modes
$N$-body simulation.
\item Move each particle in the small=modes simulation to its location at
time $a'(\xv(a)+\dvec_L)$ as defined by equation~\ref{eq:growthmatching},
interpolating between snapshots when necessary.
\end{enumerate}
The resulting particle positions give a close approximation to the particle positions
that would be obtained by running an $N$-body simulation with the Fourier modes
in $\delta_L$ present in the initial conditions (with the amplitudes suitably scaled
according to the linear growth function).
The two operations
of adding Zel'dovich displacements and perturbing
the time where the particle positions are evaluated do not commute
when the density field has evolved nonlinearly from the initial
conditions. To decide on how to order these operations we imagine
an approximation to an $N$-body simulation where the initial conditions
are first evolved by adding Zel'dovich displacements to the particles
and then the density is scaled by the linear growth function.
Reversing these operations, with the Zel'dovich move applied last,
would be less analogous to the way the particle positions
actually evolve in a simulation. In this latter scenario, the
particles would first gravitationally collapse at rates determined by the local
mean matter density and then at late times would undergo displacements to
expand or compress regions in over- or overdense volumes. There is
nothing physical about this ordering of operations.
We have confirmed with our test simulations
that the mode resampling does not work well unless the Zel'dovich move
is the first step of the mode-resampling algorithm.
\subsection{Algorithm implementation}
\label{sec:implementation}
The algorithm in Section~\ref{sec:modeaddition} describes how to add large-scale
Fourier modes to a simulation that had zero large-scale power in the
initial conditions. Ideally we would like to first remove the existing
large scale modes from a previously run simulation and then apply
the algorithm in Section~\ref{sec:resampling}. We will discuss why this
is challenging in Section~\ref{sec:subtractingpower},
but first we describe details of the implementation
of the mode addition algorithm.
To generate a simulation without large-scale power (what we call
a ``small-modes'' simulation), we generated
Gaussian Fourier mode realizations on a grid only for those modes, set all modes
with $|\kv| < k_{\rm thresh}$ to zero, and then performed the reverse
Fourier transform to get the initial density field that is used for the
first particle displacement step. We then evolve the simulation normally,
but saving 43 snapshots equally spaced in $\ln(a)$ over the range
$0.1 \lesssim a \lesssim 5.4$ so the
particle positions at perturbed times can be accurately determined
by interpolating between snapshots.
For testing purposes we ran another simulation using the
same initial conditions, but before the large-scale modes were set to
zero (called an ``all-modes'' simulation).
The particle positions at $a=1$ in the all-modes simulation
give our benchmark for testing the reconstruction of the density field
when the large-scale modes that were deleted from the initial conditions
are added back into the small-modes simulation at $a=1$.
Note that all the phases in the initial conditions were identical for both
the all-modes and small-modes simulations.
Unless otherwise specified, all our simulations were in box
$1\, h^{-1}{\rm Gpc}$ on a side with $256^{3}$ particles and were
evolved using a lean version of the Gadget-2 code~\citep{gadget2b,gadget2a} from Gaussian
initial conditions at redshfit 20.
Because gravitational evolution couples small-scale Fourier modes to the
largest modes in the simulation box, the nonlinear perturbation growth rate
in the small-modes simulation (i.e., with large-scale modes missing)
is smaller than that in the all-modes simulation,
as illustrated with the power spectrum comparison in Figure~\ref{fg:psgrowthrates}.
\begin{figure}
\centerline{
\includegraphics[scale=0.5]{ps_takahashi_allz.pdf}
}
\caption{\label{fg:psgrowthrates}Matter power spectra for several snapshots
of the simulations with scale factor increasing from bottom to top. The solid lines
are for the all-modes simulation, dashed lines are for the small-modes
simulation and dotted lines show the linear theory prediction. Notice the
divergence of the nonlinear all-modes and small-modes spectra as
the simulation proceeds. }
\end{figure}
The solid lines show the power spectra of select snapshots in the all-modes simulation
while the dashed lines show power spectra of the small-modes simulation at the
same times (recall that the power spectra are equal in the initial conditions).
At late times, the all-modes power spectrum has a significantly larger amplitude
on small scales than the small-modes power spectrum. A first test of our
mode-addition algorithm will be to correct for this deficit of nonlinear power.
In order for the growth-matching condition in equation~(\ref{eq:growthmatching}) to
provide a useful means for correcting for missing large-scale power,
the errors in matching the growth rate by perturbing the matter density
must be much smaller than the change in the growth rate due to the ommission
of large-scale power. We tested this condition explicitly by running
two simulations with different values of
$\Omega_m$ and we plot their power spectra at matched linear growth in
Figure~\ref{fg:omegam_match}. The power spectra in the two cosmologies
match to $<0.1$\% for wavenumbers $k < 0.2$~$h/$Mpc, which is much
less variation than between the all-modes and small-modes power spectra.
This is why matching the linear
growth is a good way to mimic the change in the local matter density.
\begin{figure}
\centerline{
\includegraphics[scale=0.45]{varying_omegam_spectra.pdf}
}
\caption{\label{fg:omegam_match}Power spectra relative to the all-modes spectra
from simulations with
different values of $\Omega_m$ plotted at matched values of the
linear growth. For comparison, the power spectra for our fiducial
cosmology with and without large-scale modes are also shown. The
variation in the spectra with different $\Omega_m$ is only apparent
for $k \gtrsim 0.5$~$h$Mpc$^{-1}$ and even there it is much less
than the difference between the spectra with large-scale modes removed.}
\end{figure}
We note that \citet{zheng02} have performed similar tests and showed that when the
linear theory power spectrum is fixed (as we have assumed),
several clustering statistics besides just the nonlinear power spectrum can
be easily related between models with different $\Omega_m$ and $\sigma_8$.
Next, we tested the various steps of the mode addition algorithm
from Section~\ref{sec:modeaddition} by matching the particle positions between
the all-modes and the small-modes simulations
by adding the large-scale modes from the all-modes initial conditions
back into the small-modes simulation at $a=1$ (after scaling the large-scale
modes by the linear growth function).
Figure~\ref{fg:dotplots} illustrates our comparison for a section of the
$1 (h^{-1}{\rm Gpc})^{3}$ simulation volume. The black points show the
particle positions in the all-modes simulation while red shows the positions
in the small-modes simulation.
It is readily apparent to the eye that both the time-perturbation
and Zel'dovich moves are necessary to obtain a good match to the particle
positions in the all-modes simulation.
\begin{figure*}
\centerline{
\includegraphics[scale=0.5]{dotplot_takahashi_algorithmcomparison.png}
}
\caption{\label{fg:dotplots} Particle positions in a section of our 1~Gpc$^{3}/h^{3}$ simulation box (of width 50~Mpc$/h$). The red points denote the particle positions in a simulation with the large--scale Fourier amplitudes omitted from the initial conditions as described in the text. The black points denote positions in a simulation with all Fourier modes included, but that is otherwise identical to the simulation denoted by the red points. The 4 panels show various components of our algorithm for replacing the missing large--scale power. The top left panel has no correction applied to the red points, the top right has the red particles shifted to their positions at different times so that the linear growth rate mimics the perturbation to the local matter density from the added large--scale modes. The bottom left panel shows a Zel'dovich move applied to the red points, where the Zel'dovich displacements are derived from the large--scale modes only. Finally the bottom right panel shows our full algorithm combining the time-perturbation and Zel'dovich displacements.
The axis labels are in $h^{-1}$Mpc.}
\end{figure*}
The Zel'dovich move alone does a better job matching the particle positions in the largest over-densities than
the time-perturbation alone, but leaves significant position offsets compared to the final algorithm combining
both perturbations.
\subsection{Subtracting large-scale power}
\label{sec:subtractingpower}
The algorithm we have demonstrated requires that an $N$-body simulation be run with
initial conditions that have the large-scale Fourier mode amplitudes set to zero and
with a relatively large number of late time snapshots saved that can be used for
interpolating particle positions. Ideally, we would like to have an algorithm that
can be applied to existing $N$-body simulations. Such an algorithm would first
require a method for subtracting the large-scale modes already present in the
simulation. Our procedure for adding large-scale modes should work in reverse,
but with the new growth-matching condition,
\begin{equation}\label{eq:inversegrowthmatching}
D(a',\Omega_m (1 + \delta_{L}(\xv, a)) = D(a,\Omega_m).
\end{equation}
We arrive at this condition by modeling regions where $\delta_{L}$ is overdense (overdense)
as regions that have evolved without large-scale power but with a larger (smaller) local value
of $\Omega_{m}$.
We then find the time
$a'$ where a simulation with density $\Omega_m(1+\delta_{L}(\xv,a))$ and no large-scale
power would have the same linear growth as the original simulation with density
$\Omega_m$ at time $a$.
However, removing the effect of large-scale under-densities can require evolving those overdense
regions far into the future to match the growth on the right-hand side of equation~(\ref{eq:inversegrowthmatching}).
For $\Lambda$CDM, the local linear growth in large under-densities can freeze and will never
match the right-hand side in equation~(\ref{eq:inversegrowthmatching}) even if evolved infinitely into the future.
The growth matching condition in equation~(\ref{eq:inversegrowthmatching}) and the problem
with under-densities is illustrated in Figure~\ref{fg:growthtableinverse}.
\begin{figure}
\centerline{
\includegraphics[scale=0.45]{growth_tables_inverse.pdf}
}
\caption{\label{fg:growthtableinverse}Scale factor versus matter density at matched linear growth for mode subtraction. Left panel: The scale factor versus the linear growth function normalized to one at the time of the initial
conditions ($a_{IC}=1/21$) for different values of $\Omega_{m}$ over the range $[0.01,0.5]$. Each
black line represents a different value of $\Omega_{m}$ starting with $\Omega_{m}=0.01$ for the left-most
line and increasing in steps of 0.05.
The solid vertical red line shows the value of the normalized growth function at $a=1$ and
the target mass density $\Omega_{m}^{0}=0.238$.
Right panel: The growth matching condition for removing large-scale modes in equation~(\ref{eq:inversegrowthmatching})
amounts to finding the value of the scale factor in the left panel where the red vertical line intersects the black lines
for each $\Omega_{m}$ value. The line of intersection points is plotted as the black line in the right panel.
The vertical dotted lines show the range of $\Omega_{m}^{0}(1\pm4.3\sigma(\delta_{L}))$ (as in
Figure~\ref{fg:scalefactor_vs_omegam}). The scale factor value matching the upper bound in
the perturbation to $\Omega_{m}$ is shown by the dotted horizontal line. There is no
scale factor value that can match the linear growth at the lower bound of the perturbed $\Omega_{m}$,
rendering the mode subtraction impossible for the overdense regions of $\delta_{L}$.}
\end{figure}
For our fiducial cosmology, $L_{\rm box}=1\,h^{-1}{\rm Gpc}$, and $k_{\rm thresh}=8 k_{F}$
all regions in $\delta_{L}$ that have density smaller than the mean density minus $\sim2\sigma$
will be impossible to correct using the growth matching condition of equation~(\ref{eq:inversegrowthmatching}).
In the bottom panel of Figure~\ref{fg:avslbox} we show the maximum $a'$ derived from
equation~(\ref{eq:inversegrowthmatching}) as a function of simulation box size assuming
$k_{\rm thresh} / k_{F}=8$ and that the maximum variations in $\delta_{L}(\xv,a)$ are $4.3\sigma$.
For cubic simulation boxes with sides less than $\sim 1.8~h^{-1}{\rm Gpc}$ there is
no solution for $a'$. For larger box sizes the maximum $a'$ quickly decreases from $\sim6$ to
very close to one. So for simulation box sizes of several $h^{-1}$Gpc,
removing large-scale modes may be feasible with minimal extra computational effort.
However, again for computational expediency, we have limited our considerations in this
paper to our $1~h^{-1}$Gpc simulations and therefore do not consider the
removal of large-scale modes further.
\section{Validation tests}
\label{sec:validation}
We present further tests in this section to assess the accuracy of our
method for introducing
large-scale modes in $N$-body simulations. We are primarily interested
in the matter power spectrum and power spectrum covariance matrix
and investigate estimators for these in turn.
\subsection{Recovering the nonlinear power spectrum}
In Figure~\ref{fg:addpowersnapshots} we demonstrate the reconstruction of the
nonlinear power spectrum from the small-modes simulation.
Using the all-modes simulation as our benchmark, we used the
Fourier modes from the all-modes initial conditions with $|\kv| < k_{\rm thresh}$
to determine the Zel'dovich moves and perturbed times for adjusting
the small-modes simulation (so if our algorithm worked perfectly the
power spectra should be perfectly matched).
In addition to the $\Lambda$CDM simulations described in
Section~\ref{sec:implementation},
we ran analogous all-modes and small-modes simulations for an
SCDM cosmology (i.e., with $\Omega_{m}=1$).
Because the SCDM model has no dark energy, the linear growth
rate is completely determined by the matter density, and
our growth matching condition in equation~(\ref{eq:growthmatching})
should be an especially good approximation for the effect of
large-scale Fourier modes on the gravitational growth rate.
\begin{figure*}
\centerline{
\includegraphics[scale=0.53]{addPowerSnapshots_scdm_ng512_scaled.pdf}
\includegraphics[scale=0.53]{addPowerSnapshots_takahashi_ng512_scaled.pdf}
}
\caption{\label{fg:addpowersnapshots}Power spectra relative to the all-modes spectra
demonstrating the
accuracy with which large-scale power can be added to a simulation
without any large-scale modes. Left: SCDM, right: $\Lambda$CDM.
The large-scale modes ($\delta_{L}$) added to the simulations without
large-scale power (small-modes) have the same phases as in the simulation
run with all modes present (shown by the gray horizontal lines). The vertical
dashed line marks the Nyquist frequency of the density fields.}
\end{figure*}
Figure~\ref{fg:addpowersnapshots} shows that the fractional errors in the reconstructed
power spectra are indeed slightly smaller for the SCDM simulation, but the errors are much
less than 1\% for both cosmologies. Due in part to the cloud-in-cell deconvolution
(which over-corrects the power on small scales),
the fractional errors begin to increase significantly at about $0.5\,k_{\rm Nyquist}$.
Note that the Zel'dovich correction alone,
shown by dotted blue lines in Figure~\ref{fg:addpowersnapshots},
does a very poor job of correcting the small-scale power spectrum
for the added large-scale modes. This seems to be in contrast to the
lower left panel of Figure~\ref{fg:dotplots} where the Zel'dovich correction
alone appears to correct the particle positions rather well. We now see that
the eye notices the agreement of large-scale structures more easily than the
disagreement of small-scale structures in the dot plots of
Figure~\ref{fg:dotplots}.
\subsection{Power spectrum covariance matrix}
The gravitational growth of perturbations increases the small-scale
power spectrum variance and induces significant correlations between
power spectrum bands~\citep{scoccimarro99, meiksin99}.
In this Section we test how well our mode
addition algorithm can reproduce these effects.
By perturbing a single small-modes simulation with independent
Gaussian realizations of the large-scale modes, we can
generate many realizations of the density field.
These are not independent realizations because the small-scale modes are
initially the same for each resampling.
We use the covariance matrix estimated in~\citet{takahashi09} from 5000 $N$-body
simulations to test our sample covariance estimates from simulations
with resampled large-scale modes.
\subsubsection{Sample covariance matrix estimator}
First we define the estimator for the covariance matrix that we use
for our comparisons.
Let $\delta_{\kv_i}$ denote the amplitude at grid point $\kv_i$ of the Discret
Fourier Transform (DFT) of the density field from an $N$-body simulation
measured on a 3D uniform periodic grid.
Then, the standard estimator for the power spectrum of the mode
amplitudes is obtained by averaging $\left|\delta_{\kv_i}\right|^2$ over a
spherical shell of constant radius $\left|\kv_i\right|$,
\begin{equation}
\phat_q = \frac{V}{\Nkq} \sum_{\kv_i, \left|\kv_i\right|=k_q} \left|\delta_{\kv_i}\right|^2,
\end{equation}
where $\Nki$ is the number of modes that fit in the shell.
We follow~\citet{takahashi09} and set this shell thickness to $0.01~h{\rm Mpc}^{-1}$.
If we generate $N_r$ realizations of the density field (e.g. by
running $N_r$ identical $N$-body simulations except with different
random number seeds to generate the initial conditions), the mean
power spectrum estimator for all the realizations is:
\begin{equation}\label{eq:meanpowerspectrumestimator}
\bar{P}_q = \frac{1}{N_r}\sum_{a=1}^{N_r}\,\phat_q^{a}
\end{equation}
The estimator for the covariance of the power spectrum from all the realizations is then:
\begin{equation}
\hat{C}_{ij} = \frac{1}{N_r}\sum_{a=1}^{N_r}
\left( \phat_i - \bar{P}_i \right)\,
\left( \phat_j - \bar{P}_j \right).
\end{equation}
\citet{meiksin99}, \citet{norberg08}, and \citet{takahashi09} found that
at least several hundred
realizations of the matter density field are required to accurately
estimate the sample covariance matrix.
This means that hundreds (or thousands in some cases) of $N$-body
simulations must be run with different random number seeds in order
to obtain a power spectrum covariance matrix estimate that is sufficiently accurate
for cosmological parameter estimation.
\subsubsection{Multiple power spectra from a single simulation box}
Because running hundreds or thousands of $N$-body simulations is
often prohibitively expensive,
previous studies have attempted to estimate the power spectrum
covariance from a single $N$-body simulation by applying window
functions to the gridded density field to sub-sample the simulation volume.
\citet{hamilton06} found that because the windows are convolved with
the Fourier modes of the periodic simulation box, the resulting
power spectrum covariance matrix estimates are significantly biased
with respect to the result obtained from an ensemble of periodic simulations.
Furthermore this bias cannot be simply accounted for by deconvolving the
window functions from the power spectrum estimates, but is a result of
higher order connected correlations entering the covariance matrix estimator,
which \citet{hamilton06} call ``beat-coupling.''
Although direct sub-sampling of the simulation volume has been shown not
to reproduce the covariance from an ensemble of simulations, the idea that different
regions of a large simulation volume provide statistically independent information about the
small-scale covariance is sound if we accept the ergodic hypothesis and assume our
simulation is big enough.
To avoid the convolution in Fourier space imposed by the configuration-space
windows of~\citet{hamilton06}, we attempted to sub-sample the Fourier modes that
enter the shell-averaged power spectrum estimator in equation~(\ref{eq:meanpowerspectrumestimator}).
For shells with large wavenumbers there are thousands of modes used to estimate
the power spectrum. The nonlinear power spectrum covariance depends on the connected
four-point correlations of the Fourier modes, but with thousands of modes on a uniform grid
there must be many redundant four-point configurations that could yield independent
samples of the (shell-averaged) power spectrum covariance (after appropriate scaling
by the number of modes used to estimate the mean power spectrum).
The enumeration and counting of the four-point configurations in each shell is complicated,
so we instead divided the surface of each wavenumber shell
into uniform area HEALPix\citep{healpix} pixels
and then selected only one wavevector within each pixel to obtain
a power spectrum estimate. For a given number of pixels covering the spherical shell, the
minimum multiplicity of wavevectors in a pixel determines the number of pseudo-independent
power spectrum estimates we can derive with this method.
That is, if all pixels have at least $n$ wavevectors within their boundaries,
then we can generate at most
$n$ pseudo-independent realizations of the power spectrum estimates.
We defined smaller (and therefore more) pixels for shells of increasing wavevector magnitude.
This ensured that the multiplicity of each pixel was greater than one for a wide range
in wavevector magnitudes.
We applied this
sub-sampling to a small-modes simulation with 500 large-scale mode resamplings.
The resulting power spectrum estimates and correlation coefficients of the
estimated power spectrum covariance are shown in Figure~\ref{fg:distributedpowerspectrum}.
\begin{figure
\centerline{
\includegraphics[scale=0.3]{distributed_powerspectrum.pdf}
}
\centerline{
\includegraphics[scale=0.3]{distributed_correlationcoefficients.pdf}
}
\caption{\label{fg:distributedpowerspectrum}Top: Power spectrum estimates from
using only a sub-sample of the Fourier modes in each wavevector shell.
The power spectra are normalized by the power spectrum of the all-modes simulation
at $a=1$, so the wiggles at small $k$ are indicating the sample variance in the
all-modes simulation.
Each power spectrum estimate from sub-sampled $k$-shells
is shown by the grey lines in the top panel.
The bottom panel of the top figure shows the number of Fourier modes used
to estimate the shell-averaged power spectrum. The large jumps are where
we changed the number (and size) of the pixels used to sub-sample each $k$-shell.
Bottom: Correlation coefficients of the power spectrum estimates (solid, black) compared
to the result from \citet{takahashi09} (dashed, red).}
\end{figure}
While the mean power spectrum is successfully recovered with this algorithm,
the power spectrum covariance estimator has nearly zero off-diagonal terms, in stark
contrast to the result from \citet{takahashi09} (shown in red, dashed lines) for wavenumbers
$\gtrsim 0.1$~$h$Mpc$^{-1}$.
One explanation for this could be that our method of sub-sampling modes in the
wavevector shells is not actually selecting representative samples of all the
four-point configurations in each shell, which would be fixed by a more precise sub-sampling
procedure.
Because pursuing this algorithm further would significantly increase the computational
complexity while limiting the range of applications, we instead turned to
using multiple small-modes simulations with large-scale mode resampling
to further explore the covariance matrix estimates as described in the next Section.
\subsubsection{Covariance matrix estimates from mode-resampling}
\label{sec:covestimation}
The power spectrum variance and
correlation coefficients estimated from 500 resamplings of the large-scale modes
in a single small-modes simulation are shown by the blue squares in
Figure~\ref{fg:covcomparison}.
Note that we use the same binning in wavenumber in all the cases shown as~\citet{takahashi09}
in order to make an accurate comparison.
\begin{figure*}
\centerline{
\includegraphics[scale=0.5]{var_norm_base.pdf}
\includegraphics[scale=0.5]{cor_base.pdf}
}
\caption{\label{fg:covcomparison} Left: Power spectrum variance
normalized to that for a Gaussian density field with the same linear power.
Right: Select correlation coefficients of the power spectrum covariance.
Each panel corresponds to a row of the matrix of correlation coefficients, with
the row labeled by the annotated $k$ value in each panel (which is also where the
correlation coefficients are all equal to one each panel).
The vertical dashed lines show the value of $k_{\rm thresh} \equiv 8 k_F$. The colors and
point styles are the same for both panels. The black circles are values from
\citet{takahashi09}. The red diamonds are derived from the sample covariance estimate
from 500 resamplings of each of 20 small-modes simulations. The blue squares are the
estimates using 500 resamplings of only 1 small-modes simulation. Finally, the green
$x$'s are the estimates when using 20 simulations without any resamplings (i.e., the
sample covariance estimate that would be derived without knowledge of the algorithm
in this paper.)}
\end{figure*}
While the small-scale variance estimate is boosted beyond the Gaussian prediction,
it is systematically biased low relative to the results of~\citet{takahashi09}.
At intermediate wavenumbers just larger than $k_{\rm thresh}$,
the estimated variance is only $\sim20\%$ of the true value. This is not
too surprising because it is at these scales where our algorithm for adding
large-scale modes is expected to be most inaccurate.
This is because our approximation that longer wavelength modes simply change
the effective local matter density must be wrong when the longer-wavelengths
are nearly equal to the smaller wavelengths to be perturbed. The modes just
larger and smaller than $k_{\rm thresh}$ may both contribute to coherent structures in the
mass density distribution so that resampling only some of the Fourier modes in these
structures would severely underestimate the true variance on these scales.
Also, we should note that the modes with wavenumbers just smaller than $k_{\rm thresh}$
are likely not strictly Gaussian distributed in the all-modes simulation, so that their true
distribution and coupling to small modes is not properly represented.
The right panel in Figure~\ref{fg:covcomparison} shows four rows of the matrix of
estimated correlation coefficients of the power spectrum.
It is significant that the mode resampling generates nonzero correlation coefficients
between the large and small scales (blue crosses -- lower panels especially),
which are identically zero in the
unperturbed small-modes simulation. However, the correlation coefficients
are again biased low by multiplicative factors of two or more.
We have checked that the biases in the estimated power spectrum variance
and correlation coefficients do not change when using more than 500
large-scale mode realizations. Instead we speculate that this bias is real
and indicates that (at least for fixed $k_{\rm thresh}$) the mode resampling algorithm
generates realizations of the matter density from only a subset of the full
nonlinear density probability distribution. This makes sense intuitively
if we realize that the mode resampling as implemented here can never
generate a new set of halos in the simulation box, but
only moves and merges or de-merges the
halos already present in the small-modes simulation.
Several previous
investigations have shown that the phase correlations in the Fourier
modes of the density field in a dark
matter simulation are dominated by the masses and positions of the most massive
halos~\citep{coles03,neyrinck07}. It makes sense then that the power spectrum variance on
small-scales would also be quite sensitive to the particular realization of the most massive halos.
However, with a larger simulation volume the covariance estimates might be less
dependent on the particular realization of the massive halos because many different
halo configurations would be present.
If our mode resampling algorithm is simply sampling a subset of the density probability
distribution, then the bias in the power spectrum variance and correlation coefficients
should decrease if we use more than one simulation.
To check this hypothesis we ran 19 more small-modes simulations (for 20 total)
with different random number seeds in the initial conditions and
added 500 resamplings of the large-scale modes to each simulation (for a total of 10000
realizations of the density field).
The sample covariance estimates improve considerably as shown by the
red diamonds in Figure~\ref{fg:covcomparison}. The biases in both the variance and
the correlation coefficients are much smaller over all scales. In addition,
the extremely low dip in the variance estimates at intermediate scales when using only
one simulation has disappeared entirely with 20 simulations.
Finally, we show an estimate of the covariance matrix using 20 simulations without any
resampling of large-scale modes with the green crosses in Figure~\ref{fg:covcomparison}. With such a
small number of simulations the statistical noise dominates in the estimator. Comparing with the
estimator with 500 resamplings of each of the $N$-body simulations shows how much our
mode-addition algorithm reduces the statistical errors.
\subsubsection{Convergence rate of covariance estimates}
Although with 20 small-modes simulations the bias in the power spectrum covariance
estimate is much smaller than with only one simulation, a small bias is still noticeable
in Figure~\ref{fg:covcomparison}. In Figure~\ref{fg:convergence} we attempt to
quantify the improvement in the covariance matrix estimates as more
small-simulations are added by plotting the r.m.s. dispersion in elements of the
power spectrum variance relative to the result from~\citet{takahashi09}.
\begin{figure*}
\centerline{
\includegraphics[scale=0.46]{dispconv.pdf}
\includegraphics[scale=0.46]{dispconv_nresamp.pdf}
}
\caption{\label{fg:convergence} Left: Dispersion in components of the power spectrum covariance matrix estimates as functions of the number of $N$-body simulations used to
estimate the covariance. The dispersion is measured relative to the result from~\citet{takahashi09}
using 5000 simulations. Each $N$-body simulation has 500 resamplings of the large-scale modes, so a
point on the plot has (number of $N$-body simualtions) $\times$ 500 total power spectra used to estimate the
covariance. The boxes indicate the medians (central black lines) and quartiles (top and bottom of the
boxes) for the dispersion of the diagonal covariance elements within the wavenumber bins
$0.295\le k/{\rm (h/Mpc)}\le 0.395$. The dashed black line indicates the fit to the dispersion found
in~\citet{takahashi09}. The red line is a fit to the medians of our dispersion measurements.
Right: Normalization of the fit to $\sigma^2_{\rm cov}$ versus $N_r$ as a function of the number
of mode resamplings per $N$-body simulation. That is, we plot the fit parameter
$A\equiv 2/\left(\sigma^2_{\rm cov}N_{r}^b\right)$ just as for the red solid line in the left panel,
but using between 1 and 500 mode resamplings per $N$-body simulation. (We fit a different $b$ parameter
for each value on the abcissa, but do not show these values.)
The error bars denote the
95\% confidence intervals on the nonlinear least-squares parameter estimate. The dashed grey line
shows the expected scaling if each mode resampling was equivalent to running a new $N$-body simulation.
}
\end{figure*}
We show the distribution in the variance estimates for wavenumber bins with centers in
$0.295 < k/({\rm Mpc}/h) < 0.395$ with the boxes in the left panel of Figure~\ref{fg:convergence}.
The horizontal
lines in the centers of the boxes denote the median of the variance estimates and the box top and
bottom denote the first and third quartiles.
We have selected only high-$k$ modes to plot in Figure~\ref{fg:convergence} because the
variance estimate in this range has a roughly constant bias as seen in Figure~\ref{fg:covcomparison}
and we are primarily interested in the accuracy of the covariance estimates on small scales where
the deviation from the Gaussian model is most severe. We show how the normalization of the fit
to the covariance dispersion changes with the number of mode resamplings per $N$-body
simulation in the right panel of Figure~\ref{fg:convergence}. The reduction in the number of
$N$-body simulations needed to give a certain error in the estimate covariance flattens
around 100 resamplings per simulation.
\citet{takahashi09} performed a similar analysis and found that the dispersion in their
sample variance estimators was well fit by $\sigma = 2 /N_{r}$, where $N_{r}$ is the
number of $N$-body simulations (or realizations) used to compute the estimator (shown
as the dashed black line in Figure~\ref{fg:convergence}).
We find a nonlinear least squares fit to our variance estimates of $2/(8.5 N_{r}^{0.8})$,
which is shown by the solid red line in Figure~\ref{fg:convergence}. From this we conclude
that the sample variance estimator using resampled large-scale modes can achieve
similar accuracy as the standard sample variance estimator but with $\sim 8$ times
fewer $N$-body simulations. However, it is difficult to extrapolate this convergence rate
far beyond the 20 simulations we have actually performed because small changes in
our choice of wavenumber bins or in our least-squares fit can cause large
extrapolation errors.
\section{Discussion and Conclusions}
\label{sec:conclusions}
We have demonstrated an algorithm to resample the large-scale Fourier
modes of the density
in an $N$-body simulation that successfully reproduces the nonlinear power spectrum
and significantly reduces the number of simulations required to estimate
the power spectrum covariance matrix. We expect our algorithm to aid
the calculation of uncertainties for observational probes of the matter power spectrum
such as weak lensing and galaxy clustering. Because we can quickly provide multiple
realizations of the density field, it is possible to estimate uncertainties after applying
appropriate survey windows to the density, which can have significant impacts
on the estimated power spectrum covariances. This could be a distinct advantage over
approaches that either precompute the power spectrum covariance without
considering the survey window~\citep{semboloni07, takahashi09, sato09},
or use approximate or analytic methods that
may not include the survey windows correctly~\citep[e.g.][]{crocce10}.
We have no reason to
believe the extension to redshift space will pose any particular challenges.
However, note that we have not yet tested the perturbation of the particle velocities so our
results are limited to real-space statistics for now.
Our algorithm uses both Zel'dovich displacements as well as time-perturbations
to match the effective local linear growth in over- or overdense regions.
It is intriguing that neither of these operations alone is sufficient to
reconstruct even the nonlinear power spectrum. Evaluating the time-perturbation
at the particle positions prior to the Zel'dovich move gives a bad result,
indicating that the early-time movement of the particles is significant in
determining the later growth of structures. Any extensions of our algorithm
will likely have to consider both of these components carefully.
Because applying the time-perturbation to remove large-scale modes from an
existing simulation generally requires many simulation snapshots far into the
future, we found it onerous (or even impossible if resampling too many modes in
a box that is too small)
to use existing simulations for mode-resampling.
However, if the simulation volume is at least several Gpc on a side, then
according to Figure~\ref{fg:growthtableinverse} it may be possible to subtract a
limited number of large-scale modes with a feasible amount of computation.
For simulations volumes smaller than many Gpc on a side,
we advocate running a new simulation with large-scale modes removed
in the initial conditions, which can then have new large-scale modes added
directly in later snapshots using our algorithm. Note that even for adding
new large-scale modes the time-perturbation requires many snapshots
closely spaced in scale factor both in the past and future of the desired
time (but the simulation does not have to be run as far into the future as is
needed for removing large-scale modes). This means that application of
our algorithm for, e.g. parameter estimation from a galaxy survey, will
require special-purpose $N$-body simulations that will save significant
computation time for the error analysis.
Our algorithm has other applications such as
including scaling
of halo masses and concentrations so that mock galaxy catalogues could be
constructed from each of the resampled density fields. In our view, this is the
only viable method to obtain robust uncertainties for estimating cosmological
parameters from galaxy clustering and BAOs. Our algorithm may even
provide a feasible method to go beyond the work of~\citet{sefusatti06} for
estimating the covariance of higher-order correlations.
For parameter estimation
with weak lensing, our resampled density fields could be directly input
to existing ray-tracing pipelines. However, one would need resampled
densities for each of the lens planes (which could extend to $z\sim3$ for
upcoming survey requirements). This would require saving even more
closely spaced snapshots of the small-modes simulation(s).
The sample covariance matrix estimated from a large number of resamplings
of a single simulation has nonzero correlations between large and small scales
that are introduced by our mode addition algorithm. But, the covariance elements
are biased low relative to the benchmark result from~\citet{takahashi09}, which
used 5000 $N$-body simulations. A possible explanation for this bias is that our
method of resampling large-scale modes samples only a subset of the full
probability distribution for the nonlinear density field. We have shown that the bias
decreases monotonically as a function of the number of
$N$-body simulations used to estimate the covariance, which is consistent
with this explanation. If our mode-resampling algorithm were perfect at adding
large-scale fluctuations, the bias in the covariance that we find
would indicate that the distribution of the
small-scale nonlinear density field is not entirely determined by the coupling
with large-scale modes. Rather, the phases of the small-scale Fourier modes in
the initial conditions would be important as well, which contradicts the conclusion
of~\citet{little91} who found little impact on the final
nonlinear density when the small-scale phases were randomized in the
initial conditions. This could be because the eye notices mostly the
large-scale structures in a dotplot of the particle positions
but there can be significant small-scale deviations that are less noticeable
(as shown by the ``Zel'dovich only'' lines in Figure~\ref{fg:addpowersnapshots} versus the
``Zel'dovich only" lower left panel of Figure~\ref{fg:dotplots}).
One area for future investigation is to study the dependence of the bias in
the power spectrum sample covariance (or other measures of the distribution of
the nonlinear density) with $k_{\rm thresh}$. This would quantify the relative importance
of different large-scale modes on the nonlinear density distribution. This could
be a promising way to better understand the ``loss'' of information about
cosmology in the matter power spectrum on
translinear scales~\citep{rimes06, neyrinck06, neyrinck07}.
One could also quantify on which scales (if any) the density in a large simulation
becomes ergodic (so that the distribution of the density in sub-volumes of
the simulation is equivalent to the distribution of the density field in an ensemble
of large simulations). The fact that our small-scale covariance estimates have
such a large bias for $k_{\rm thresh}\sim 0.5\,h/$Mpc (and in a single simulation) suggests
that assuming ergodicity on these scales is a fallacy when defining estimators
of, at least some, summary statistics of the density.
Another application of our algorithm could be for tiling smaller simulation
volumes to obtain the (properly correlated) density field in volumes that are
too large to simulate, as originally proposed by~\citet{cole97}. This would probably
be most effective if we run several of our ``small-modes'' simulations and then
tile them together by resampling the modes both larger and smaller than the
fundamental frequency of the simulation volume. A related application is re-running
resimulation studies (that take a sub-volume of high-resolution $N$-body simulation and
add additional physics) with different large-scale mode realizations. This could be
a relatively fast way to obtain different realizations of the resimulations with only
one high-resolution $N$-body simulation at hand.
Finally, we hope to apply our algorithm for obtaining estimates of both the mean
power spectrum and its covariance as needed in simulation emulator
frameworks~\citep{heitmann09, schneider11}. Because we can estimate the
power spectrum covariance with an easily achievable number of simulations,
it will now be feasible to build an emulator for the cosmological-parameter
dependence of the covariance~\citep{schneider08}. Note that this paper is
complementary to~\citet{angulo10} who described how to rescale the
outputs of a single $N$-body simulation to mimic the outputs with
a different cosmology, but who did not consider the time-perturbations we
use here. In combination, our algorithm and that of~\citet{angulo10}
may serve to drastically reduce the number of simulations needed to construct
simulation emulators for cosmological analysis.
\section*{Acknowledgments}
We thank Adrian Jenkins for extensive technical advice on Gadget-2
and setup of initial conditions, David Weinberg for pointing us to his
earlier related work, Alex Szalay for explanations of the growth of
correlations in the Fourier phases of the density field, Yanchuan Cai
for advice on applying perturbation theory to (an ultimately failed attempt to)
add large-scale modes to our simulations and
Mark Neyrinck, Bhuvnesh Jain, and Ravi Sheth for
useful conversations.
SMC acknowledges the support of a Leverhulme Research Fellowship.
Some of the calculations for this paper were performed on the ICC
Cosmology Machine, which is part of the DiRAC Facility jointly funded by STFC,
the Large Facilities Capital Fund of BIS, and Durham University.
This work performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344.
\bibliographystyle{apj}
|
\section{Introduction}
\medskip
Random matrices of various ensembles find numerous applications in
several fields of statistical physics.
In the general class of non-hermitian random matrices an
important role is played by the {\sl Ginibre ensemble} \cite{Gi65}.
A matrix $G$ of size $N$ of such an ensemble consists of $N^2$
independent random complex numbers, drawn according to the
Gaussian distribution with zero mean and a fixed variance \cite{Me04,Fo10}.
Such matrices are used to describe non-unitary dynamics of
chaotic systems and open quantum systems \cite{Ha10}.
This ensemble of random matrices can also be used
to analyze the human EEG data \cite{Se03},
for telecommunication applications based on the scattering
of electromagnetic waves on random obstacles \cite{TV04},
or in mathematical finances
to describe correlation matrices of various stocks \cite{BP,BJN03}.
The spectrum of a non-Hermitian matrix $G$
belongs to the complex plane.
Spectral density of random matrices of the suitably normalized
Ginibre ensemble is described by the Girko circular law \cite{Gi84},
as in the limit $N \to \infty$ it covers uniformly the unit disk.
The random matrix $W=GG^{\dagger}$, called a {\sl Wishart matrix},
is positive. Hence its eigenvalues $\lambda_i$, $i=1,\dots, N$,
are real and non-negative. Introducing a rescaled eigenvalue, $x=N\lambda$
one can show
that in the limit of the large matrix size the spectral density $P(x)$
converges to the {\sl Marchenko--Pastur} (MP) distribution \cite{MP67}.
In general, products of random matrices are a subject of an intensive research
for many years \cite{CPV93}.
Recent studies on products of Ginibre matrices
concern multiplicative diffusion processes \cite{GJJN03},
correlation matrices used in macroeconomic time series \cite{BLMM07},
a random matrix approach to quantum chromodynamics \cite{Os04}
and lattice gauge field theories \cite{LNW08}.
Properties of the complex spectra of products of random Ginibre matrices
were recently analyzed in \cite{BJW10}.
It is also interesting to study singular values of a product
of $s$ independent Ginibre matrices, $X=G_1\cdots G_s$.
Note that a squared singular value of the product $X$ equals
the corresponding eigenvalue of the Wishart like matrix $W=XX^{\dagger}$.
For $s=2$, positive random matrices of the form
$W_2=G_1G_2 (G_1 G_2)^{\dagger}$ found their
applications in finances \cite{BLMM07}.
Matrices of the form $W_s=G_1 \cdots G_s (G_1 \cdots G_s)^{\dagger}$
for an arbitrary $s$ were used to describe random quantum states
associated with certain graphs \cite{CNZ10}
and quantum states obtained by orthogonal
measurements in a product of
maximally entangled bases \cite{ZPNC10}.
The corresponding asymptotic level density distribution $P_s(x)$ is
called {\sl Fuss-Catalan distribution} of order $s$,
since its moments are given \cite{armstrong,Ml10,banica-etal}
by the Fuss-Catalan numbers \cite{GKP,Ko08},
(also called Fuss \footnote{When Leonard Euler, after an eye operation in 1772,
became almost completely blind, he asked Daniel
Bernoulli in Basel to send a young assistant, well trained in mathematics,
to him in St.~Petersburg.
It was Nikolaus Fu{\ss} who arrived in St.~Petersburg in May 1773 \cite{Oz}.}
numbers \cite{PS00}). Strangely enough, the Fuss-Catalan numbers
generalize the Catalan numbers, although the work of Fu{\ss} \cite{Fu791}
was done much earlier then the contribution of Catalan \cite{Ca838}.
The Catalan number can be defined
as a number of different bracketing of a product of $n+1$ numbers,
or the number of possible $n$ folding of a
map which contains $n+1$ pages in a row \cite{Ko68}.
The Fuss--Catalan distribution describes asymptotically statistics
of singular values of the $s$--th power of
random Ginibre matrices.
This result obtained recently by Alexeev et al. \cite{AGT10}
was derived by estimating the moments of the
distribution of squared singular values of a power $G^s$
of a random matrix and showing that these moments
converge asymptotically to the Fuss--Catalan numbers.
This is true under rather weak assumptions:
all entries of the matrix $G$ are independent random variables
characterized by the zero mean, variance set to unity
and finite fourth moment.
The Fuss--Catalan distribution can be considered as a generalization
of the MP distribution, which is obtained for $s=1$.
Moreover, the distribution $P_s(x)$
belongs to the class of free Meixner measures \cite{bozejko-bryc},
and in terms of free probability theory
it appears as the {\sl free}
multiplicative convolution product of $s$ copies of the
MP distribution \cite{banica-etal,BG09},
which is written as $P_s(x)=[P_1(x)]^{\boxtimes s}$.
An explicit form in the case $s=2$ was derived in \cite{PS01} in context of
construction of generalized coherent states from combinatorial sequences.
The spectral distribution of $P_s(x)$ for a product of an arbitrary number
of $s$ random Ginibre matrices was recently analyzed by Burda et al. \cite{BJLNS10}
also in the general case of rectangular matrices. The distribution was expressed
as a solution of a polynomial equation and it was conjectured that
the finite size effects can be described by a simple multiplicative correction.
Another recent work of Liu at al. \cite{LSW10}
provides an integral representation of the distribution $P_s(x)$
derived in the case of $s$ square matrices of size $N$, which is assumed to be large.
However, these recent contributions do not provide an explicit
form of the distribution $P_s(x)$.
The aim of this note is to derive exact and explicit
formulae for the Fuss-Catalan distribution $P_s(x)$,
which can be represented as a combination of $s$
hypergeometric functions.
The derivation is presented in Sec. II, while
some auxiliary information on special functions and the
proof of positivity of $P_s(x)$
are provided in Appendices A and B, respectively.
In section III we discuss a certain two-parameters
generalization of Fuss-Catalan numbers. As these numbers
quantify generalized Raney sequences \cite{GKP},
the corresponding probability measures $W_{p,r}(x)$ will be called
{\sl Raney distributions}. As special cases they include
Marchenko--Pastur distribution, Fuss--Catalan distributions
and the Wigner semicircle law. We
find exact expressions for Raney distributions
corresponding to integer parameter values in the general case
and provide explicit formulae in the simplest cases
of small values of the integer parameters $1\le r\le p$.
\section{Fuss--Catalan distributions}
For any integer number $s$ one can use the binomial symbol
to define a sequence of integers
denoted by $FC_s(n)$,
\begin{equation}
FC_s(n) := \frac{1}{sn+1}\binom{sn+n}{n} \ .
\label{FCmom}
\end{equation}
Here $n=0,1,\dots$, while $s=1,2,\dots$, and
these numbers are called the {\sl Fuss--Catalan} numbers of order $s$ \cite{GKP}.
We enumerate some of these sequences below for $n=0,\dots,7$:
\begin{eqnarray}
FC_1(n) & = & 1,1,2,5,14,42,132,427,\dots \nonumber \\
FC_2(n) & = & 1,1,3,12,55,273,1428,7752,\dots \nonumber \\
FC_3(n) & = & 1,1,4,22,140,969,7084,53820,\dots \nonumber \\
FC_4(n) & = & 1,1,5,35,285,2530,23751,231880,\dots \; . \nonumber
\label{FCseqen}
\end{eqnarray}
The above sequences are contained in the
Online Encyclopedia of Integer Sequences (OEIS) \cite{Sloane}
under the labels (A000108), (A001724), (A002293)
and (A002294), respectively.
These sequences can be considered as a generalization
of the sequence $FC_1(n)$,
which consists of {\sl Catalan numbers},
$C(n)=\frac{1}{n+1} \binom{2n}{n}$.
We are going to show that for any given $s$
there exists a density distribution $P_s(x)$, which satisfies
\begin{equation}
\int_0^{K_s} x^n P_s(x) dx = FC_s(n), \ \ n=0,1,\dots
\label{densmom}
\end{equation}
where
\begin{equation}
K_s:=(s+1)^{s+1}/s^s \; .
\label{Ks}
\end{equation}
In other words we are looking for a positive density $P_s(x)$
which satisfies the above infinite system of equations.
As the density turns out to be defined in a finite segment
$[0,K_s]$, the solution of this
{\sl Hausdorff moment problem} \cite{Ak65}
associated with the Fuss-Catalan numbers is unique.
An explicit proof of positivity of $P_s(x)$
is provided in Appendix B.
We employ the method of the inverse Mellin transform,
which was previously used to construct explicit solutions
of the Hausdorff moment problem \cite{KPS01,GPHDS}
and to derive explicit form of the L{\'e}vy--stable distributions \cite{PG10}.
The {\sl Mellin transform} ${\cal M}$ of a function $f(x)$ and its inverse
${\cal M}^{-1}$ are defined by a pair of equations
\begin{equation}
f^*(\sigma):= \; {\cal M}[f(x);\sigma] = \int_0^{\infty}
x^{\sigma-1} f(x) dx
\label{mellin1}
\end{equation}
and
\begin{equation}
f(x):= \; {\cal M}^{-1}[f^*(\sigma);x]
= \frac{1}{2\pi i} \int_{c-i\infty}^{c+i\infty} x^{-\sigma} f^* (\sigma) d\sigma,
\label{mellin2}
\end{equation}
with complex $\sigma$.
In (\ref{mellin2}) this variable is integrated over a vertical line in
the complex plane \cite{FGD95}.
For discussion concerning the role of $c$, see (\cite{FGD95,PBM98,Lu69}).
Therefore the solution of Eq. (\ref{densmom}) can be obtained by extending
integer variable $n$ to complex $\sigma$ by a substitution $n\to \sigma -1$.
The desired form of the distribution $P_s(x)$ can be formally written as an inverse
Mellin transform,
\begin{equation}
P_s(x) \ = \ {\cal M}^{-1}[FC_s(\sigma);x] \ .
\label{mellin3}
\end{equation}
To find such a transform we will bring
the Fuss-Catalan numbers into a more suitable form.
Representing the binomial symbol in (\ref{FCmom})
by the ratio of Euler's gamma function one obtains
\begin{equation}
FC_s(\sigma) =
\frac{\Gamma\bigl[ (s+1)\bigl(\sigma -\frac{s}{s+1}\bigr) \bigr] }
{\Gamma \bigl[ s \bigl( \sigma - \frac{s-2}{s} \bigr) \bigr]
\; \Gamma (\sigma) } \ .
\label{FSbis}
\end{equation}
After applying twice the Gauss--Legendre formula (\ref{multeuler})
for multiplication of the argument of
the gamma function one arrives at
\begin{eqnarray}
FC_s(\sigma) &= & \frac{1}{\sqrt{2\pi}}
\Bigl[\frac{(s+1)^{s+1}}{s^s}\Bigr]^{\sigma} \times
\nonumber \\
& \times & \;
\frac{s^{s-3/2}}{(s+1)^{s+1/2}} \;
\Bigl[ \prod_{j=0}^{s-1}
\frac{\Gamma\bigl(\sigma+ \frac{j-s}{s+1}\bigr)}
{\Gamma\bigl(\sigma+ \frac{2+j-s}{s}\bigr)}
\Bigr] .
\label{FStri}
\end{eqnarray}
Obtaining the above form of the FC numbers,
in which a ratio of products of the gamma functions of a shifted argument
appears, is a key step of our reasoning.
It allows us to represent the inverse Mellin transform of Eq. (\ref{FStri})
as a certain special function.
To see this recall that the {\sl Meijer $G$--function}
of the argument $z$
can be defined by the inverse Mellin transform \cite{PBM98}:
\begin{eqnarray}
\label{Meijer}
&& G_{p,q}^{m,n} \Bigl( z \; \bigl|\; {\stackrel{\scriptstyle
\alpha_1 \cdots \alpha_p}
{\scriptstyle \beta_1 \cdots \beta_q}} \Bigr) =
\\
&& = {\cal M}^{-1}
\Bigl[
\frac{\prod_{j=1}^{m}\Gamma(\beta_j+\sigma) \prod_{j=1}^{n}\Gamma(1-\alpha_j-\sigma)}
{\prod_{j=m+1}^{q}\Gamma(1-\beta_j-\sigma)
\prod_{j=n+1}^{p}\Gamma(\alpha_j+\sigma)}; \; z
\Bigr] . \nonumber
\end{eqnarray}
This definition involves four lists of parameters,
which can be represented by $p$ complex numbers $\alpha_j$ and
other $q$ complex numbers $\beta_j$.
Integers numbers $p$ and $q$ can be equal zero and it s is assumed
that $0\le m\le q$ and $0 \le n \le p$, so that possibly empty products
in this form are taken to be equal to unity.
A detailed description of the integration contours of the Mellin transform
(\ref{Meijer}), general properties of the Meijer functions and its special
cases can be found in \cite{PBM98}.
\begin{figure}[hbp]
\centering
\includegraphics[width=0.32\textwidth]{FC12B.eps}
\caption{ (Color online) Marchenko--Pastur distribution $P_1(x)$
compared with the Fuss--Catalan distribution $P_2(x)$.
The singularity at $x\to 0$
is of the type $P_s(x)\sim x^{-s/(s+1)}$.}
\label{fig1}
\end{figure}
\begin{figure}[htbp]
\centering
\includegraphics[width=0.36\textwidth]{FC3456aB.eps}
\caption{ (Color online) Fuss--Catalan distributions $P_s(x)$
plotted for parameter $s$ equal to $3,4,5$ and $6$ are supported on the interval
$[0,(s+1)^{s+1}/s^s]$. To show behaviour at the right
side of the support
the figure is plotted for $x \ge 5$.
}
\label{fig2}
\end{figure}
Direct comparison of expression (\ref{FStri}) for the Fuss Catalan numbers
and the Mellin transform as the Meijer $G$--function (\ref{Meijer})
allows us to represent the Fuss-Catalan distribution $P_s(x)$
by a Meijer $G$--function,
\begin{equation}
P_s(x) = \frac{1}{\sqrt{2\pi}} \frac{s^{s-3/2}}{(s+1)^{s+1/2}} \;
G_{s,s}^{s,0} \Bigl( z \; \bigl|\; {\stackrel{\scriptstyle
\alpha_1 \cdots \alpha_p}
{\scriptstyle \beta_1 \cdots \beta_q}} \Bigr)
\label{FCdist1}
\end{equation}
of the argument $z=x s^s (s+1)^{-(s+1)}$.
Looking at the range of the parameter $j$ in (\ref{FStri})
we see that the numbers of parameters of the Meijer $G$--function
have to be set to $n=0$, $p=s$ and $m=q=s$.
Hence this function involves $2s$ parameters, which read
$\alpha_j=(1+j-s)/s$ and $\beta_j=(j-1-s)/(s+1)$ for $j=1,\dots, s$.
As there are only two products of gamma functions
in (\ref{FStri}), in contrast to four in
(\ref{Meijer}), which is equivalent to setting $n=0$ and $m=q$,
a further simplification of the above formula is possible.
In this very case the Meijer $G$--function
can be written as a combination of hypergeometric functions
of the same argument $z$ -- see Eq. (5.2.11) p. 146 of \cite{Lu69}.
Let ${}_pF_q\Bigl( \bigl[ \{a_j\}_{j=1}^p \bigr] , \;
\bigl[ \{b_j\}_{j=1}^q \bigr] ; \; x\Bigr)$
denote the hypergeometric function \cite{MOT} of the type ${}_pF_q$
with $p$ 'upper' parameters $a_j$ and $q$ 'lower' parameters $b_j$
of the argument $x$. The symbol
$\{a_i\}_{i=1}^r $ represents the list of $r$ elements, $a_1,\dots a_r$.
Then formula (\ref{FCdist1}) for the Fuss--Catalan
distribution can be rewritten as
\begin{widetext}
\begin{equation}
\label{eq:FCs}
P_s(x) =
\sum_{k=1}^s \Lambda_{k,s}\; x^{\frac{k}{s+1}-1}\;
{}_sF_{s-1}\Bigl( \Bigl[ \Bigl\{1-\frac{1+j}{s} +\frac{k}{s+1} \Bigr\}_{j=1}^s \Bigr] , \;
\Bigl[ \Bigl\{1+\frac{k-j}{s+1} \Bigr\}_{j=1}^{k-1} ,
\Bigl\{1+\frac{k-j}{s+1} \Bigr\}_{j=k+1}^{s} \Bigr] ; \;
\frac{s^s}{(s+1)^{s+1}} x
\Bigr) \ .
\end{equation}
where the coefficients $\Lambda_{k,s}$ read for $k=1,2, \dots, s$
\begin{equation}
\label{eq:FCs2}
\Lambda_{k,s} := s^{-3/2} \sqrt{\frac{s+1}{2\pi}}
\Bigl(\frac{s^{s/(s+1)}}{s+1}\Bigr)^{k} \
\frac{\Bigl[ \prod_{j=1}^{k-1} \Gamma\bigl(\frac{j-k}{s+1}\bigr) \Bigr]
\Bigl[ \prod_{j=k+1}^{s} \Gamma\bigl(\frac{j-k}{s+1}\bigr) \Bigr] }
{ \prod_{j=1}^{s} \Gamma\bigl( \frac{j+1}{s} - \frac{k}{s+1}\bigr) } \ .
\end{equation}
\end{widetext}
This formula, along with Eq.(\ref{raney2}) below, constitutes
the key result of the present note.
It gives an exact result for the Fuss--Catalan distribution $P_s(x)$
for an arbitrary natural $s$.
The FC distribution describes the density of squared singular
values of a product of $s$ independent square Ginibre matrices
in the limit of large matrix size.
The convergence conditions of the hypergeometric series
${}_sF_{s-1}$ immediately yield the support of $P_s(x)$
which is equal to $[0,(s+1)^{s+1}/s^{s}]$.
For small values of $x$ the distribution behaves as $x^{-s/(s+1)}$.
It is conforting to see that in the simplest case $s=1$ the above
complicated form reduces indeed to
the Marchenko--Pastur distribution,
\begin{equation}
P_1(x) = \frac{1}{\pi \sqrt{x}}
\; {}_1F_0 \Bigl( [-\frac{1}{2}], [ \ ] ; \; \frac{1}{4}x \Bigr)
=\frac{\sqrt{1-x/4}}{\pi \sqrt{x}}\; .
\label{pi1bis}
\end{equation}
Furthermore, the distribution $P_2(x)$, shown in Fig. \ref{fig1},
\begin{eqnarray}
P_2(x) =&& \frac{\sqrt{3}}{2\pi x^{2/3}}
\;{}_2F_1 \Bigl( \bigl[-\frac{1}{6},\frac{1}{3}\bigr], \;
\bigl[\frac{2}{3} \bigr] ; \; \frac{4}{27}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
- \frac{\sqrt{3}}{6\pi x^{1/3}}
\;{}_2F_1 \Bigl( \bigl[\frac{1}{6},\frac{2}{3}\bigr] , \;
\bigl[\frac{4}{3}\bigr]; \; \frac{4}{27}x \Bigr)
\label{p21bis}
\end{eqnarray}
is equivalent to the form,
\begin{equation}
\!\!
P_2(x) = \frac{\sqrt[3]{2} \sqrt{3}}{12 \pi} \;
\frac{\bigl[\sqrt[3]{2} \left(27 + 3\sqrt{81-12x} \right)^{\frac{2}{3}} -
6\sqrt[3]{x}\bigr] } {x^{\frac{2}{3}}
\left(27 + 3\sqrt{81-12x} \right)^{\frac{1}{3}}},
\label{pi2}
\end{equation}
valid for $ x \in [0,27/4]$ and
obtained first in \cite{PS01} in context of
construction of generalized coherent states from combinatorial sequences.
The distribution $P_3(x)$, plotted in Fig. \ref{fig2},
is given by a sum of three terms,
\begin{eqnarray}
P_3(x) = \!\!
&& \frac{1}{\sqrt{2} \pi x^{3/4}}
\;{}_3F_2 \Bigl( \bigl[-\frac{1}{12},\frac{1}{4}, \frac{7}{12}\bigr], \;
\bigl[\frac{1}{2},\frac{3}{4}\bigr]; \; \frac{27}{256}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
\!\!
- \frac{1}{4 \pi x^{1/2}}
\;{}_3F_2 \Bigl( \bigl[ \frac{1}{6},\frac{1}{2}, \frac{5}{6}\bigr], \;
\bigl[ \frac{3}{4},\frac{5}{4}\bigr]; \; \frac{27}{256}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
\!\!
- \frac{\sqrt{2}}{64 \pi x^{1/4}}
\;{}_3F_2 \Bigl( \bigl[\frac{5}{12},\frac{3}{4}, \frac{13}{12}\bigl], \;
\bigl[\frac{5}{4},\frac{3}{2}\bigr]; \; \frac{27}{256}x \Bigr) .
\label{pi3}
\end{eqnarray}
In Fig. 2 we present the Fuss--Catalan distributions
$P_s(x)$ for $s=3,4,5$ and $6$.
\section{Raney distributions}
\label{sec_Raney}
The Fuss--Catalan numbers $FC_s(n)$ defined in (\ref{FCmom}) can be
considered as a special cases of a larger family of sequences,
\begin{equation}
R_{p,r}(n) := \frac{r}{pn+r}\binom{pn+r}{n}
\label{Ran1}
\end{equation}
defined for $n=0,1,\dots$.
Here $p$ and $r$ are treated as integer parameters,
$p\ge 2$, $r=1,2,\dots$ .
Setting $r=1$ and $p=s+1$ we have
$\frac{1}{np+1}\binom{np+1}{n}=\frac{1}{n(p-1)+1}\binom{np}{n}$,
so the numbers $R_{s+1,1}(n)$ are equal to $FC_s(n)$.
Further relations involving the sequences $R_{p,r}(n)$ are
\begin{equation}
R_{p+1,p+1}(n) = FC_p(n+1),
\label{Ranrel1}
\end{equation}
and
\begin{equation}
R_{p,p}(n) = R_{p,1}(n+1) ,
\label{Ranrel2}
\end{equation}
which can be verified directly from their definitions (\ref{FCmom}) and (\ref{Ran1}).
The Raney lemma \cite{Ra60} implies
that the number of the Raney sequences of order $p$ and length $pn+1$,
for which all partial sums are positive,
is given by the Fuss--Catalan numbers $FC_{p-1}(n)=R_{p,1}(n)$.
Furthermore, as the number of positive generalized Raney sequences
is equal to $R_{p,r}(n)$ \cite{GKP}
we will refer to $R_{p,r}(n)$ defined in (\ref{Ran1})
as {\sl Raney numbers}. These numbers
appear as coefficients in a generalized binomial series \cite{GKP}.
Some representative examples of sequences $R_{p,r}(n)$,
for $n=0,1,\dots,7$ are quoted below
together with their OEIS labels \cite{Sloane},
\begin{eqnarray}
R_{4,2}(n) & = & 1,2,9,52,340,2394,17710,135720, ... (A069271) \nonumber \\
R_{5,2}(n) & = & 1,2,11,80,665,5980,56637,556512, ... (A118969), \nonumber
\label{ranseq1}
\end{eqnarray}
whereas two following sequences are not represented in OEIS:
\begin{eqnarray}
R_{4,5}(n) & = & 1,5,30,200,1425,10626,81900,647280, ... \; ,\nonumber \\
R_{6,3}(n) & = & 1,3,21,190,1950,21576,250971,3025308,... \; . \nonumber
\label{ranseq2}
\end{eqnarray}
\medskip
In a recent work M{\l}otkowski \cite{Ml10}
has shown that the sequence (\ref{Ran1})
describes moments of a probability measure $\mu_{p,r}$ with a compact support
contained in $[0,\infty)$,
if point $(r,p)$ which determines parameters of the Raney numbers
belongs to the set $\Sigma$ defined by inequalities
$p\ge 0$ and $0 < r \le p$.
Note that the point $(1,1)$ implies a constant sequence of moments,
$R_{1,1}(n)=1$, which represents a singular, Dirac delta measure,
$\mu_{1,1}=\delta(x-1)$.
In the case the measure $\mu_{p,r}$
is represented by a density we will denote it by $W_{p,r}(x)$.
Setting $r$ to one one gets the Fuss--Catalan numbers,
which implies that $W_{2,1}(x)$ represents the Marchenko--Pastur distribution,
while $W_{s+1,1}(x)$ reduces to the
Fuss--Catalan probability density $P_s(x)$.
In general parameters $p$ and $r$ can be taken to be real
and then the moments of the measure are expressed by the gamma functions,
\begin{equation}
\int x^n \mu_{p,r}(x) dx =
\frac{r}{np+r}
\frac{\Gamma(np+r+1)} {\Gamma(n+1) \Gamma(np+r-n+1)} ,
\label{Ran2}
\end{equation}
where the integration covers entire support of the measure $\mu_{p,r}$.
For $1 \le r \le p$ the distribution $W_{p,r}(x)$ is a positive function,
see \cite{Ml10} and Appendix B.
The corresponding distribution $W_{p,r}(x)$ can be written implicitly
by its $S$-transform, which allowed M{\l}otkowski
to establish relations between various Raney distributions \cite{Ml10},
listed in Appendix C.
For a precise definition of the $S$--transform
(or the free multiplicative transform) see Eq. 4.9 in \cite{Ml10}.
In spite of these concrete results
an explicit form of the Raney distribution $W_{p,r}(x)$
has not appeared in the literature so far.
Making use of the inverse Mellin transform and the Meijer function
we can generalize results of the previous
section and obtain explicit expressions
for the Raney distributions $W_{p,r}(x)$,
which correspond to integer values of the parameters $p$ and $r$.
Repeating steps analogous to Eqs.(\ref{FSbis}) -- (\ref{FCdist1})
we can represent distribution
$W_{p,r}(x)$ in terms of the Meijer $G$--function.
The explicit expression generalizing Eq.(\ref{FCdist1})
reads
\begin{equation}
W_{p,r}(x) = \frac{r}{\sqrt{2 \pi}} \frac{p^{r-p-1/2} } {(p-1)^{r-p+3/2}}
\; G_{p,p}^{p,0} \Bigl( z \; \bigl|\; {\stackrel{\scriptstyle
\alpha_1 \cdots \alpha_p}
{\scriptstyle \beta_1 \cdots \beta_q}} \Bigr) ,
\label{raney}
\end{equation}
where the argument of the function is $z=x(p-1)^{p-1}/p^p$,
while its parameters are $\alpha_1=0$, $\alpha_j=(r-p+j)/(p-1)$ for
$j=2,\dots, p$
and $\beta_j=(r-p-1+j)/p$ for $j=1,\dots, p$.
In analogy to (\ref{eq:FCs}) this relation can be
represented by the following sum consisting of of $p$ terms:
\begin{widetext}
\begin{equation}
\!\!
W_{p,r}(x) =
\sum_{j=1}^{p} \Omega (p,r;j) x^{\frac{r-1+j}{p}-1}\;
{}_pF_{p-1}\Bigl( \Bigl[
1+\beta_j,
\Bigl\{1+\beta_j-\alpha_{i}\Bigr\}_{i=2}^p \Bigr] , \;
\Bigl[\Bigl\{1+\frac{j-i_2}{p} \Bigr\}_{i_2=1}^{j-1},
\Bigl\{1+\frac{j-i_3}{p} \Bigr\}_{i_3=j+1}^{p} \Bigr] ; \;
\frac{(p-1)^{p-1}}{p^p} x \Bigr) \; ,
\label{raney2}
\end{equation}
where ${}_pF_{p-1}$ is the hypergeometric function and the
numerical coefficients $\Omega(p,r;j)$, for $j=1,2, \dots, p$
read
\begin{equation}
\label{eq:raney3}
\Omega(p,r;j) := \frac{r}{\sqrt{2\pi}}
\frac{p^{r-p-1/2}}{(p-1)^{r-p+3/2}}
\Bigl(\frac{(p-1)^{p-1}}{p^p}\Bigr)^{\frac{r-p-1+j}{p}} \;
\frac{1}{ \Gamma \bigl( \frac{p-r+1-j}{p} \bigr)}\;
\frac{\Bigl[ \prod_{i_1=1}^{j-1} \Gamma\bigl(\frac{i_i-j}{p}\bigr) \Bigr]
\Bigl[ \prod_{i_2=1}^{p-j} \Gamma\bigl(\frac{i_2}{p}\bigr) \Bigr] }
{ \prod_{i_3=2}^{p} \Gamma\bigl(\frac{r-p+i_3}{p-1} - \frac{r-p-1+j}{p}\bigr) } \ .
\end{equation}
\end{widetext}
Convergence properties of the hypergeometric function
imply that the Raney distribution $W_{p,r}(x)$ for integer values of
its parameters is supported in the interval $[0,K_{p-1}]$,
where $K_s$ is given in Eq. (\ref{Ks}).
Formula (\ref{raney2}) implies that for $p>r$ the distribution $W_{p,r}(x)$
diplays a singularity for small $x$ of the type $x^{-(p-r)/p}$.
For $p=r$ the 'diagonal' Raney distributions behave for small arguments
as $W_{p,p}(x) \sim x^{1/p}$.
It is helpful to add some clarifying remarks concerning the
key formula (\ref{raney2}).
We draw attention to the fact that some simplifications will
always appear for two reasons:
a) firstly, one parameter form the 'upper' list of the parameters
will be always equal to a parameter from the 'lower' list.
Consider, for instance, the case $p=4$ and $r=2$. Then the value
of the first parameter in the 'upper' list becomes $1+(j-3)/4$
and it cancels with the value of $i_3=3$ of the first sequence
of the 'lower' parameters. One can demonstrate that
a similar cancellation effect takes place for any pair $(p,r)$.
Therefore, the hypergeometric function ${}_pF_{p-1}$ in (\ref{raney2})
effectively reduces to ${}_{p-1}F_{p-2}$.
b) secondly, we see from Eq.(\ref{eq:raney3})
that the coefficient $\Omega(p,r;j)$ vanishes for $j=p+1-r$, due to
the presence of the first gamma function in denominator. Therefore the
sum in Eq. (\ref{raney2}) involves $(p-1)$ terms with different hypergeometric functions
${}_{p-1}F_{p-2}$. Then it can be explicitly verified that from the equality
between the numbers $R_{s+1,1}(n)=FC_s(n)$, the corresponding equality
between the probability distributions
\begin{equation}
W_{s+1,1}(x) \ = \ P_s(x)
\label{raney6}
\end{equation}
follows. In particular the following identity between the coefficients holds,
$\Omega(s+1,1;j)=\Lambda_{j,s}$, whose demonstration
is rather tedious and will not be reproduced here.
With the two provisos explained under items a) and b) above,
Eq.(\ref{raney2}) will be used to obtain explicit forms
of distributions $W_{p,r}(x)$ for small values of $r$ and $p$.
In particular, the case $W_{2,2}(x)$
reduces to the celebrated semi-circular law \cite{Me04}:
\begin{equation}
\!
W_{2,2}(x) = \frac{\sqrt{x}}{\pi}
\; {}_1F_0 \Bigl([-\frac{1}{2}], [ \ ]; \; \frac{x}{4} \Bigr)
= \frac{1}{2\pi}
\sqrt{x(4-x)} .
\label{R22}
\end{equation}
The above semicircle is centered at $x=2$,
while the Wigner semicircle centered at $x=0$
is used in random matrix theory
to describe the asymptotic level density of
random hermitian matrices form Gaussian ensembles \cite{Me04,Fo10}.
The Raney distributions $W_{2,r}(x)$ are plotted in Fig. \ref{fip2}
for $r=1$ and $r=2$. For comparison we have also plotted the function
$W_{2,3}(x)$ furnished by Eq. (\ref{raney2}). According to the results of M{\l}otkowski
\cite{Ml10} and our Appendix B
this function is not positive in its domain,
so it does not represent a probability distribution.
Here we obtain it explicitly.
\begin{figure}[htbp]
\centering
\includegraphics[width=0.35\textwidth]{RaneyFig1B.eps}
\caption{ (Color online)
Raney distributions $W_{2,r}(x)$ with values of the parameter $r$
labeling each curve. For $r=1$ it reduces to Marchenko-Pastur distribution $P_1(x)$,
while a semicircle law is obtained for $r=2$. For $r=3$ the function
represented by dashed line is not positive, which is implied by $p<r$.
}
\label{fip2}
\end{figure}
It is easy to observe that the above semicircle distribution is related
with the Marchenko-Pastur distributions $P_1(x)$
by a relation $W_{2,2}(x)=xP_1(x)$. This is a special case
of a more general relation involving the "diagonal"
Raney distributions with $r=p$ and Fuss-Catalan distribution,
\begin{equation}
W_{p,p}(x) = x\; W_{p,1}(x) =x P_{p-1}(x).
\label{Rpp}
\end{equation}
This result, established first in \cite{Ml10},
follows also naturally from Eqs.(\ref{raney}) and (\ref{raney2}).
Thus for $p=3$ one has
\begin{eqnarray}
W_{3,1}(x) = P_2(x), \ \ {\rm and} \ \
W_{3,3}(x) = x \; P_2(x) ,
\label{W33}
\end{eqnarray}
where the Fuss-Catalan distribution $P_2(x)$ is given in
(\ref{p21bis}). Due to an equivalent expression (\ref{pi2})
the distribution $W_{3,3}(x)$ can also be expressed in terms of elementary
functions.
The intermediate Raney distribution, corresponding to $r=2$,
\begin{eqnarray}
W_{3,2}(x) =&& \frac{\sqrt{3}} {2\pi x^{1/3}}
\;{}_2F_1 \Bigl( \bigl[ -\frac{1}{3},\frac{1}{6}\bigr], \;
\bigl[ \frac{1}{3}\bigr] ; \; \frac{4}{27}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
- \frac{\sqrt{3} x^{1/3}}{18\pi}
\;{}_2F_1 \Bigl( \bigl[\frac{1}{3},\frac{5}{6}\bigr] , \;
\bigl[\frac{5}{3}\bigr] ; \; \frac{4}{27}x \Bigr)
\label{W32}
\end{eqnarray}
is shown in Fig. \ref{fip3}.
In a close analogy to Eq. (\ref{pi2})
this distribution enjoys
a similar representation in terms of elementary functions,
\begin{equation}
\!\!
W_{3,2}(x) = \frac{\sqrt{3} \sqrt[3]{2}}{36 \pi} \;
\frac{\bigl[ \left(27 + 3\sqrt{81-12x} \right)^{\frac{4}{3}} -
18 \sqrt[3]{2} x^{\frac{2}{3}} \bigr] }
{ x^{\frac{1}{3}}
\left(27 + 3\sqrt{81-12x} \right)^{\frac{2}{3}}},
\label{W32b}
\end{equation}
Observe that the distribution $W_{3,1}(x)$, $W_{3,2}(x)$ and $W_{3,3}(x)$
behave for small $x$ as $x^{-2/3}$, $x^{-1/3}$ and $x^{1/3}$, respectively.
\begin{figure}[htbp]
\centering
\includegraphics[width=0.35\textwidth]{RaneyFig2B.eps}
\caption{ (Color online)
As in Fig.\ref{fip2} for Raney distributions $W_{3,r}(x)$
supported in $[0,6\frac{3}{4}]$.
The case $W_{3,4}(x)$ (dashed line) is not a probability measure.
Each curve is labelled by the value of $r$.
}
\label{fip3}
\end{figure}
Let us now discuss the case $p=4$ illustrated in Fig. \ref{fip4}.
Due to relations (\ref{Rpp}) one has
\begin{eqnarray}
W_{4,1}(x) = P_3(x), \ \ {\rm and} \ \
W_{4,4}(x) = x \; P_3(x) ,
\label{W44}
\end{eqnarray}
where the Fuss-Catalan distribution $P_3(x)$ is given in
(\ref{pi3}).
Two intermediate Raney distributions have a similar form
\begin{eqnarray}
W_{4,2}(x) = \!\!
&& \frac{1}{\pi x^{1/2}}
\;{}_3F_2 \Bigl( \bigl[-\frac{1}{6},\frac{1}{6}, \frac{1}{2}\bigr], \;
\bigl[\frac{1}{4},\frac{3}{4}\bigr]; \; \frac{27}{256}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
\!\!
- \frac{\sqrt{2}}{4 \pi x^{1/4}}
\;{}_3F_2 \Bigl( \bigl[ \frac{1}{12},\frac{5}{12}, \frac{3}{4}\bigr], \;
\bigl[ \frac{1}{2},\frac{5}{4}\bigr]; \; \frac{27}{256}x \Bigr)
\ \ \ \ \ \ \ \ \nonumber \\
&&
\!\!
- \frac{\sqrt{2} x^{1/4}} {128 \pi}
\;{}_3F_2 \Bigl( \bigl[\frac{7}{12},\frac{11}{12}, \frac{5}{4}\bigl], \;
\bigl[\frac{3}{2},\frac{7}{4}\bigr]; \; \frac{27}{256}x \Bigr) ,
\label{W42}
\end{eqnarray}
and
\begin{eqnarray}
W_{4,3}(x) = \!\!
&& \frac{1}{\sqrt{2} \pi x^{1/4}}
\;{}_3F_2 \Bigl( \bigl[-\frac{1}{4},\frac{1}{12}, \frac{5}{12}\bigr], \;
\bigl[\frac{1}{4},\frac{1}{2}\bigr]; \; \frac{27}{256}x \Bigr)
\nonumber \\
&&
\!\!
- \frac{3\sqrt{2}x^{1/4} }{64 \pi}
\;{}_3F_2 \Bigl( \bigl[ \frac{1}{4},\frac{7}{12}, \frac{11}{12}\bigr], \;
\bigl[ \frac{3}{4},\frac{3}{2}\bigr]; \; \frac{27}{256}x \Bigr)
\nonumber \\
&&
\!\!
- \frac{ x^{1/2}} {32 \pi}
\;{}_3F_2 \Bigl( \bigl[\frac{1}{2},\frac{5}{6}, \frac{7}{6}\bigl], \;
\bigl[\frac{5}{4},\frac{7}{12}\bigr]; \; \frac{27}{256}x \Bigr) .
\label{W43}
\end{eqnarray}
The general formula (\ref{raney2}) allows us to obtain an explicit
form of the functions $W_{4,5}(x)$, which is not positive
and it does not represent a probability distribution,
see the dashed curve on Fig.5.
\begin{figure}[htbp]
\centering
\includegraphics[width=0.35\textwidth]{RaneyFig3B.eps}
\caption{ (Color online) As in Fig. \ref{fip2} for Raney distributions $W_{4,r}(x)$.
Dashed line corresponds to a quasi-measure
$W_{4,5}(x)$, which is not positive.
}
\label{fip4}
\end{figure}
\begin{figure}[htbp]
\centering
\includegraphics[width=0.35\textwidth]{RaneyFig4B.eps}
\caption{ (Color online) 'Diagonal' Raney distributions $W_{r,r}(x)$
form a generalization of the semicircle law $W_{2,2}(x)$.
For small $x$ they behave as $x^{1/r}$.
Value of $r$ labels each curve.
}
\label{fiprr}
\end{figure}
Figure \ref{fiprr} presents the Raney distributions $W_{p,r}(x)$
in the "diagonal" case, $r=p$.
As the distribution $W_{2,2}(x)$ represents Eq.(\ref{R22}),
the diagonal Raney distributions, $W_{r,r}(x)$,
can thus by considered as a generalization of the
semicircular law: they are defined for $x\in (0,K_{r-1})$
where $K_s=(s+1)^{s+1}/s^s$,
they are equal to zero at both ends of the domain and
they are characterized by a single maximum.
However, of $r>2$ the functions $W_{r,r}(x)$ are not symmetric
anymore, compare Fig. \ref{fiprr}.
A plot of the parameter space $(p,r)$
in which these distributions are marked
together with the Fuss-Catalan distributions
is presented in Fig. \ref{figparam}.
\begin{figure}[htbp]
\centering
\includegraphics[width=0.34\textwidth]{fc1b.eps}
\caption{Raney distributions in the parameter space $(p,r)$:
Marchenko--Pastur distribution $P_1$
is represented by the point $(2,1)$,
while Fuss-Catalan distribution $P_s$
is represented by $(*)$ at $(s+1,1)$.
The semicircle (sc) law corresponds to $(2,2)$
and the 'diagonal' Raney distributions $(+)$ to $(p,p)$.
As shown in Ref. \cite{Ml10}
any point in the shaded set $\Sigma$
corresponds to a probability measure with a compact support in $[0,\infty)$.}
\label{figparam}
\end{figure}
\section{Concluding remarks}
\label{sec_concl}
In this work we obtained an explicit form of the Fuss-Catalan distribution
$P_s(x)$. Obtained result is exact for an arbitrary $s$
and it allows for a simple use of these probability distributions.
Results derived are relevant from the point of view of statistical physics
as they describe asymptotic level density of normalized positive random matrix
of the product form $X=(G_1 \cdots G_s)(G_1 \cdots G_s)^{\dagger}$,
where $G_1, \dots ,G_s$ denote $s$ independent random matrices
from the complex Ginibre ensemble. The variable $x=N\lambda$
denotes the rescaled eigenvalue $\lambda$ of random
matrix $X$ of size $N$.
It should be emphasized here that
the Marchenko--Pastur and Fuss--Catalan distributions
describe the level density of Wishart like random matrices
in the limiting case $N\to \infty$ only.
In practice, for any fixed $N$ the finite size effects
occur. As discussed by Blaizot and Nowak \cite{BN09,BN10}
the finite $N$ effects are related with the diffraction phenomena,
while the large $N$ limit of the random matrix theory
may be compared with the geometric limit of the
wave optics or the semiclassical limit of the quantum theory.
An explicit description of finite $N$ corrections
to the spectral density of Wishart matrices
obtained from products of Ginibre matrices
is provided by Burda et al. \cite{BJLNS10,BJLNS11}.
A simple argument put forward in \cite{ZPNC10} (see Appendix B therein)
implies that the same FC distributions describe also the spectral density
of a product of $s$ random matrices taken from the {\sl real Ginibre} ensemble
\cite{LS91}.
The case of a product of two real matrices,
recently studied in context of quantum chromodynamics \cite{APS10},
is also important in applications in econophysics,
where one uses a product of two real correlation matrices \cite{BLMM07}.
The Catalan numbers are a special case of a more general,
one parameter family of Fuss-Catalan numbers, which from a subset
of two-parameter family of Raney numbers. In the same way,
the Marchenko--Pastur distribution $P_1(x)$ is a special case
of the Fuss-Catalan distributions, which in turn belong to the
two parameter family of Raney distributions $W_{p,r}(x)$.
This wide class of probability distribution
includes e.g. the Dirac delta, $\delta(x-1)$
and the semicircle law, $W_{2,2}(x)$.
Applying the inverse Mellin transform
for integer parameter values of the parameters
we found an explicit exact representation
of the Raney distributions in terms of
the hypergeometric function.
For any $r=1,2,\dots, p$ the Raney
distribution $W_{p,r}(x)$ is supported in the interval
$[0,p^p/(p-1)^{p-1}]$. For small $x$ the
distribution behaves as
\begin{equation}
W_{p,r}(x) \ \sim \
\begin{cases} x^{-\frac{p-r}{p}} & \textrm{if \ $r < p$},\\
x^{ \frac{1}{p}} & \textrm{if \ $r= p$} .
\end{cases}
\label{singul}
\end{equation}
The Raney numbers (\ref{Ran1})
imply that the mean value of the Raney distribution
$W_{p,r}(x)$ is equal to $r$,
while the second moment reads $r(2p+r-1)/2$.
Let us conclude the paper with the following remark.
The Fuss--Catalan distribution describe statistical
properties of singular values of
products of random matrices of the Ginibre ensemble.
It is then natural to ask, whether there exist
any ensembles of random matrices, such that
their squared singular values
can be described by the Raney distributions.
\bigskip
Acknowledgments.
It is a pleasure to thank G.~Akemann, Z.~Burda,
M.~Bo{\.z}ejko, B.~Collins, K.~G{\'o}rska,
A.~Horzela, W.~M{\l}otkowski, I.~Nechita and A.~M.~Nowak
for fruitful discussions and helpful remarks.
We are thankful to both referees for useful comments.
Financial support by the Transregio-12 project
of the Deutsche Forschungsgemeinschaft and
the grant number N N202~090239 of
Polish Ministry of Science and Higher Education
is gratefully acknowledged. The authors acknowledge
support from Agence Nationale de la Recherche (Paris, France)
under program PHYSCOMB No. ANR-08-BLAN-0243-2.
|
\section{Introduction}
One of the basic equations employed to describe fluctuating
macroscopic variables is the Fokker-Planck equation (FPE)
\cite{RIS:1996}. It has found applications not only in physics,
but also in other areas such as astrophysics
\cite{NM:1998,HCDHQS:2006}, chemistry \cite{NOW:1996,LHE:1995},
biology \cite{SG:1993,FBF:2003}, finance \cite{FPR:2000}, etc. In
view of its wide applicability, various methods of finding exact
and approximate solutions of the FPE's have been developed
\cite{RIS:1996,BS:1985,MP:1996,BM:1996,HU:1989,LLPS:1985,LAN:1991}.
Most of the methods, however, are concerned only with FPE's with
time-independent diffusion and drift coefficients. Generally, it
is not easy to find solutions of FPE's with time-dependent
diffusion and drift coefficients.
One of the methods of solving FPE with time-independent diffusion
and drift coefficients is to transform the FPE into a
time-independent Schr\"odinger equation, and then solve the
eigenvalue problem of the latter \cite{RIS:1996,HS:2008}. The
transformation to the Schr\"odinger equation of a FPE eliminates
the first order spatial derivative in the FPE and creates a
Hermitian spatial differential operator. In [16] it was shown
that this transformation can also be applied to transform FPE's,
which have constant diffusion coefficients and time-dependent
drift coefficients, into time-dependent Schr\"odinger equations.
Based on this connection, a perturbative approach was developed
and used to solve this special class of the FPE \cite{HD:2008}.
The diffusion equation is known to be invariant under the scale
transformation $x\rightarrow\varepsilon x$, $t\rightarrow\varepsilon^2 t$ for any
scale $\varepsilon$. Thus one can find solutions, called the
similarity solutions, of the diffusion equation with definite
scaling behaviors by means of the so-called similarity method
\cite{ZIL:1992}. Such solutions were not considered in [16]. It is
the purpose of this work to investigate the possibility of
similarity solutions of the FPE within the perturbative framework
of [16].
\section{Perturbative approach to Fokker-Planck equations}
The FPE of the probability density $W(x,t)$ in $(1+1)$-dimension
is \cite{RIS:1996}
\begin{align}\label{PA1.1}
\frac{\partial}{\partial t}W(x,t)=\Big[-\frac{\partial}{\partial x}D^{(1)}(x,t)+\frac{\partial^2}
{\partial x^2}D^{(2)}(x,t)\Big]\,W(x,t)\;,
\end{align}
where $D^{(1)}(x,t)$ and $D^{(2)}(x,t)$ are the drift and
diffusion coefficient respectively. The drift coefficient
represents the external force acting on the particles and is
usually expressed in terms of a drift potential $U(x,t)$ according
to $D^{(1)}(x,t)=-\partial U(x,t)/\partial x$.
In this paper we will focus on FPE's with constant diffusion
coefficients $D^{(2)}(x,t)=D>0$. In this case, one can solve the
FPE by exploiting the connection between the FPE and the
Schr\"odinger equation \cite{RIS:1996,HS:2008}. Setting
\begin{align}\label{PA1.2}
\psi(x,t)=\mbox{exp}\Big\{\frac{U(x,t)}{2D}\Big\}\,W(x,t),
\end{align}
one can transfrom the FPE into the Schr\"odinger equation
\begin{align}\label{PA1.3}
\frac{\partial \psi}{\partial t}=D\frac{\partial^2 \psi}{\partial x^2}
+\Big(\frac{U''}{2}-\frac{U'^2}{4D}
+\frac{\dot{U}}{2D}\Big)\psi\;,
\end{align}
where the prime and dot denote the derivatives with respect to $x$
and $t$, respectively. Each solution of the time-dependent
Schr\"odinger equation gives the corresponding solution of the FPE
via Eq.~(\ref{PA1.2}). It is, however, generally difficult to
solve the Schr\"odinger equation exactly when the potential is
time-dependent. Various approximation schemes may have to be
employed.
The perturbative approach presented in [16] to solve the FPE
containing a small parameter in the drift potential is summarized
as follows. Suppose $U(x,t)=\sum^{\infty}_{n=0}\lambda^nU_n(x,t)$,
where $|\lambda|\ll 1$ is a small parameter, and let
$\psi(x,t)=\mbox{exp}\{S(x,t,\lambda)/D\}$. Then from
(\ref{PA1.3}), we get the equation of $S(x,t,\lambda)$,
\begin{align}
\dot{S}&=DS''+S'^2+\overline{U}\;,\label{PA1.5}\\
\overline{U}&\equiv\frac{D}{2}U''-\frac{1}{4}U'^2
+\frac{1}{2}\dot{U}\;.\label{PA1.6}
\end{align}
Substituting the series form of
$S(x,t,\lambda)=\sum^{\infty}_{n=0}\,\lambda^n S_n(x,t)$ into
(\ref{PA1.5}) and collecting terms of the same order of $\lambda$,
one arrives at a set of differential equations that determine the
functions $S_n$:
\begin{align}
\dot{S}_0&=DS_0''+S_0'^2+\Big(\frac{D}{2}U_0''-\frac{1}{4}U_0'^2
+\frac{1}{2}\dot{U}_0\Big)\;,\label{PA1.7}\\
\dot{S}_1&=DS_1''+2S_0'S_1'+\Big(\frac{D}{2}U_1''-\frac{1}{2}U_0'\,
U_1'+\frac{1}{2}\dot{U}_1\Big)\;,\label{PA1.8}\\
\dot{S}_2&=DS_2''+2S_0'S_2'+S_1'^2+\Big(\frac{D}{2}U_2''
-\frac{1}{2}U_0'U_2'-\frac{1}{4}U_1'^2
+\frac{1}{2}\dot{U}_2\Big)\;,\label{PA1.9}\\
\vdots\notag\\
\dot{S}_n&=DS_n''+\sum^{\infty}_{k=0}S_k'S'_{n-k}+\mbox{terms in
$\overline{U}$ of the order $\lambda^n$}\;,\;\;\;n\geq
0\;.\label{PA1.10}
\end{align}
With $S_0$, $S_1$, $S_2$,... solved, we will have an approximate
solution $\psi(x,t)$, and hence of $W(x,t)$.
In this paper, we shall be interested in the drift potential
$U(x,t)$ of the form $U(x,t)=\lambda\,U_1(x,t)$, with all other
$U_n=0$ for $n=0$ and $n>1$. From (\ref{PA1.2}) the probability
density is
\begin{align}\label{PA1.11}
W(x,t)=e^{\frac{1}{D}(S-\frac{U}{2})}=e^{\frac{S_0}{D}}\,e^{\frac{1}{D}(\,\sum^{\infty}_{n=1}
\lambda^nS_n-\lambda\frac{U_1}{2})}\;,
\end{align}
We assume the initial profile of $W(x,t)$ of the unperturbed case
(the diffusion case) to be the delta-function, i.e.,
$W(x,t)\rightarrow\delta(x)$ at $t=0$ as $\lambda\rightarrow 0$.
Then the solution $S_0(x,t)$ is \cite{HD:2008}
\begin{align}\label{PA1.12}
S_0(x,t)=-\frac{D}{2}\ln(4\pi Dt)-\frac{x^2}{4t}\;,
\end{align}
leading to the probability density
\begin{align}\label{PA1.13}
W_0(x,t)=e^{\frac{S_0}{D}}=\frac{1}{\sqrt{4\pi
Dt}}\,e^{-\frac{x^2}{4Dt}}\;.
\end{align}
Eq.~(\ref{PA1.13}) is the well-known solution of the diffusion
equation with the delta function as the initial profile. It is
evident that under the scale transformation
$x\rightarrow\bar{x}=\varepsilon x$,
$t\rightarrow\bar{t}=\varepsilon^2 t$, $W_0(x,t)$ scales as
$W_0(x,t)=\varepsilon W_0(\bar{x},\bar{t})$.
\section{Scaling behavior}
Motivated by the scaling form of $W_0(x,t)$ in (\ref{PA1.13}), we
would like to seek similarity solutions of the perturbative FPE.
To this end, let us first obtain the required scaling behaviors of
$W(x,t)$ and $U(x,t)$ so that Eq.~(\ref{PA1.1}) is invariant under
the scale transformation
\begin{equation}
\bar{x}=\varepsilon^a x,~~\bar{t}=\varepsilon^b t, \label{scale1}
\end{equation}
where the scaling exponents $a$ and $b$ are arbitrary real
parameters. Suppose $W(x,t)$ and $U(x,t)$ scale as
$\bar{W}(\bar{x},\bar{t})=\varepsilon^{\gamma}W(x,t)$ and
$\bar{U}(\bar{x},\bar{t})=\varepsilon^{d}U(x,t)$ with real
exponent $\gamma$ and $d$, resepctively. Then Eq.~(\ref{PA1.1})
becomes
\begin{align}
\label{sb.1}
\varepsilon^{-\gamma+b}\frac{\partial
\bar{W}}{\partial
\bar{t}}=\varepsilon^{-\gamma-d+2a}\frac{\partial } {\partial
\bar{x}}\Big(\frac{\partial \bar{U}}{\partial \bar{x}}\cdot
\bar{W}\Big)+\varepsilon^{-\gamma+2a}
D\frac{\partial^2\bar{W}}{\partial \bar{x}^2}\;.
\end{align}
For Eq.(\ref{sb.1}) to have the same form as Eq.~(\ref{PA1.1}),
one must have $\varepsilon^{-\gamma+b} =\varepsilon^{-\gamma-d+2a}
=\varepsilon^{-\gamma+2a}$. This implies that $d=0$, $b=2a$, and
$\gamma$ arbitrary. It means that when the scaling exponent of $t$
is twice that of $x$, and $U(x,t)$ is scale-invariant, then the
FPE is scale invariant, and thus admits solutions $W(x,t)$ with
arbitrary scaling exponent, which is dictated by the initial
profile $W(x,0)$. This result puts a constraint on the scaling
property of the drift potential $U(x,t)$.
Applying these general results to Eq.~(\ref{PA1.11}), we see that
since $U(x,t)=\lambda U_1(x,t)$ is scale-invariant, so are all
$S_n (x,t)~(n\geq 1)$. The scaling of $W$ is therefore solely
determined by the solution $\exp\{S_0/D\}$ of the non-perturbed
FPE. With $S_0(x,t)$ given in (\ref{PA1.12}), $W(x,t)$ scales as
$W(x,t)=\varepsilon^a W(\bar{x},\bar{t})$ (see Eq.(\ref{PA1.13})).
This is in accord with the scaling of the initial profile:
$W(x,0)=\delta(x)=\varepsilon^a \delta(\bar{x})$. Our next step is
to obtain the scale-invariant solutions for the $S_n$'s. This is
attained by the similarity method, which we describe below.
\section{Similarity method}
The similarity method is a very useful method for solving a
partial differential equation which possesses proper scaling
behavior. One advantage of the similarity method is to reduce the
partial differential equation to an ordinary differential equation
through some new independent variables (called similarity
variables), which are certain combinations of the old independent
variables. In our case, the 2nd order FPE can thus be transformed
into an ordinary differential equation which may be easier to
solve. We will illustrate the method by discussing the solution
of $S_1$ below. The method applies to equations for other $S_n$ as
well.
With $U_0(x,t)=0$ and $S_0$ given in (\ref{PA1.12}), the equation
for $S_1$ is
\begin{align}\label{E1.2}
\dot{S}_1=DS_1''-\frac{x}{t}S'_1+\frac{D}{2}U_1''+\frac{1}{2}\dot{U}_1\;.
\end{align}
Let $x$ and $t$ transform according to Eq.~(\ref{scale1}). Recall
from previous discussions that $U_1$ and $S_n$ ($n\geq 1$) are
scale-invariant, i.e., $U(x,t)=\bar{U}_1(\bar{x},\bar{t})$ and
$S_n(x,t)=\bar{S}_n(\bar{x},\bar{t})$ ($n\geq 1$). In terms of
these scaled variables, Eq.(\ref{E1.2}) becomes
\begin{align}
\varepsilon^{b}\frac{\partial \bar{S}_1}{\partial
\bar{t}}&=\varepsilon^{2a}D \frac{\partial^2 \bar{S}_1}{\partial
\bar{x}^2}-\varepsilon^{b}\frac{\bar{x}}{\bar{t}} \frac{\partial
\bar{S}_1}{\partial
\bar{x}}+\varepsilon^{2a}\frac{D}{2}\frac{\partial^2
\bar{U}_1}{\partial
\bar{x}^2}+\varepsilon^{b}\frac{1}{2}\frac{\partial
\bar{U}_1}{\partial \bar{t}}\;.\label{E1.7}
\end{align}
We demand that $\bar{S}_1$ satisfies Eq.(\ref{E1.2}) and this
requires $b=2a$. To determine the form of the similarity solution,
we have to first determine a new variable $z(x,t)$ that is
invariant under the scale transformation Eq.(\ref{scale1}). From
$b=2a$, we can choose the scale-invariant similarity variable to
be
\begin{align}\label{E1.11}
z=\frac{x}{\sqrt{t}}\;.
\end{align}
Since both $S_1(x,t)$ and $U_1(x,t)$ are scaling invariant, they
must be functions of $z$ only. Hence it is reasonable to express
$S_1(x,t)$ and $U_1(x,t)$ in terms of similarity variable $z$ as
\begin{align}\label{E1.15}
S_1(x,t)=y(z);\;\;\;U_1(x,t)=u(z)\;.
\end{align}
Eq.(\ref{E1.2}) can then be cast into an ordinary differntial
equation
\begin{align}\label{E1.22}
y''(z)-\frac{z}{2D}\,y'(z)+\frac{1}{2}\,u''(z)-\frac{z}{4D}\,u'(z)=0\;.
\end{align}
Eq.~(\ref{E1.22}) is the most general equation for $S_1$ with an
arbitrary scale-invariant potential $U(x,t)=\lambda U_1(x,t)$.
Once $y(z)$ is solved, the similarity solution $S_1(x,t)$ of
Eq.(\ref{E1.2}) is obtained by putting $z=x/\sqrt{t}$. To
illustrate the procedure, we shall solve the case of a class of
simple scale-invariant potentials.
\section{$U_1(x,t)=\mu\,x^p\,t^q$}
For definiteness we consider the simplest form that may have the
required scaling property, namely, $U_1(x,t)=\mu\,x^p\,t^q$ where
$\mu$, $p$ and $q$ are arbitrary real parameters. Requiring that
$U_1$ be scale-invariant under the transformation (\ref{scale1})
with $b=2a$, we must have $q=-p/2$. From (\ref{E1.15}), we have
\begin{align}\label{E1.25}
u(z)=U_1(x,t)=\mu\,x^p\,t^q=\mu z^p\;.
\end{align}
Eq.(\ref{E1.22}) is then expressed as
\begin{align}\label{E1.27}
y''-\frac{z}{2D}\,y'=-\frac{\mu}{2}p(p-1)\,z^{p-2}+\frac{\mu\,p}{4D}\,z^{p}\;.
\end{align}
The general solution of Eq.(\ref{E1.27}) is
\begin{align}
y(z)=\alpha_0\,\int^z\,\exp\Big\{\frac{z'^2}{4D}\Big\}\,dz'+\alpha_1-\frac{\mu}{2}\,z^p
\;.
\end{align}
In order for the probability density $W(x,t)$ to be normalizable,
we require that $y(z)$ be finite as $z\to \pm\infty$. As such,
$\alpha_0$ has to be set to zero, as the exponential term is
divergent in the domain of $z\in (-\infty,\infty)$. Thus the
acceptable solution is
\begin{align}\label{E1.31}
y(z)=\alpha_1-\frac{\mu}{2}z^p\;.
\end{align}
Putting $z=x/\sqrt{t}$ back into (\ref{E1.31}) gives the solution
of the first order perturbation equation:
\begin{align}\label{E1.32}
S_1(x,t)=\alpha_1-\frac{\mu}{2}\Big(\frac{x}{\sqrt{t}}\Big)^p\;.
\end{align}
With $S_1(x,t)$ given in (\ref{E1.32}), the equation for
$S_2(x,t)$, i.e. Eq.(\ref{PA1.9}), is
\begin{align}\label{E1.33}
\dot{S}_2&=DS_2''-\frac{x}{t}S_2'\;.
\end{align}
By the procedure described in Sect.~4, this equation can be
reduced to
\begin{align}\label{E1.34}
y''-\frac{z}{2D}y'=0\;.
\end{align}
As discussed before, the finite solution is
$y(z)=\alpha_2=\mbox{constant}$. Hence $S_2$ is given by
$S_2(x,t)=\alpha_2$.
With $S_1$ and $S_2$ given above, the equations for $S_n$ ($n\geq
3$) turn out to have the same form as that for $S_2$, i.e.,
Eq.~(\ref{E1.33}). Therefore the solutions for $S_n$ ($n\geq 3$)
are $S_n(x,t)=\alpha_n=\mbox{constant}$.
From (\ref{PA1.11}), the general solution $W(x,t)$ to FPE in the
case $U_1=\mu x^pt^q$, where $q=-p/2$, is
\begin{align}\label{E1.35}
W(x,t)&\propto\mbox{exp}\Big\{-\frac{x^2}{4Dt}-\frac{\lambda\mu}
{D}\Big(\frac{x}{\sqrt{t}}\Big)^p+\frac{1}{D}\Big(\lambda
\alpha_1+\lambda^2 \alpha_2+\lambda^3
\alpha_3+\cdots\Big)\Big\}\notag\\
&\propto\mbox{exp}\Big\{-\frac{x^2}{4Dt}-\frac{\lambda\mu}{D}
\Big(\frac{x}{\sqrt{t}}\Big)^p\Big\}\;.
\end{align}
In the last step, the constants $\alpha_1$, $\alpha_2$, $\cdots$,
are set to zero since the part exp$\{\frac{1}{D}(\lambda
\alpha_1+\lambda^2 \alpha_2+\lambda^3 \alpha_3+\cdots)\}$ can be
absorbed into the normalization constant, which however can not be
determined until the parameters $p$ and $\mu$ are given.
Furthermore, for $W(x,t)$ to be normalizable, $p$ can only take
the values $p=0,1,2$, and $p={\rm even\ integer}$ for $p>2$.
Below we shall consider two special but interesting cases of such a form of drift potentials.
\subsection{$U_1=x/\sqrt{t}$, $p=1$ case}
Let us now consider the case with $\mu=1$ and $p=1$. The FPE is
\begin{align}\label{E1.36}
\frac{\partial W}{\partial t}=D\frac{\partial^2 W}{\partial
x^2}+\frac{\lambda} {\sqrt{t}}\frac{\partial W}{\partial x}\;.
\end{align}
Its similarity solution is
\begin{align}\label{E1.37}
W(x,t)=\frac{1}{\sqrt{4\pi Dt}}\,\mbox{exp}\Big\{-\frac{1}{4Dt}\Big(x+2\lambda\sqrt{t}
\Big)^2\Big\}\;.
\end{align}
This result is the same as that in [16]. One can see from
(\ref{E1.37}) that position $x$ is shifted as time $t$ elapses.
\subsection{$U_1=x^2/2t$, $p=2$ case}
The second example is the case with $U_1(x,t)=x^2/2t$, where the
parameters taken are $\mu=1/2$ and $p=2$. From the discussions
given before, we find that the probability distribution function
\begin{align}\label{E1.38}
W(x,t)=\sqrt{\frac{1+2\lambda}{4\pi Dt}}\,\mbox{exp}\Big\{-\frac{x^2}{4Dt}(1+2\lambda)\Big\}\;,
\end{align}
is the solution to the FPE
\begin{align}\label{E1.39}
\frac{\partial W}{\partial t}=D\frac{\partial^2 W}{\partial x^2}+\frac{\lambda x}{t}\frac{\partial W}
{\partial x}+\frac{\lambda}{t}W\;.
\end{align}
In (\ref{E1.38}), one can see that the perturbation parameter
$\lambda$ controls the half width of the Gaussian distribution
$W(x,t)$.
\section{Summary}
When a partial differential equation possesses certain scaling
behavior, the so-called similarity method is of great help in
finding its solutions. One advantage of the similarity method is
to reduce the partial differential equation to an ordinary differential equation through
some new independent variables (called similarity variables),
which are certain combinations of the old independent variables.
The solutions so obtained, called similarity solutions of the
differential equations, likewise possess proper scaling forms. A
well-known example is the diffusion equation.
In this paper, we have applied the similarity method to a class of
perturbative FPE's with small time-dependent drift potentials
studied in [16]. We have presented the main ideas of the
similarity method. The method was then applied to find similarity
solutions of the FPE with certain scale-invariant drift
potentials. Our results show that similarity method can be a
useful tool to solve FPE with time-dependent drift and diffusion
coefficients. While the present work is only concerned with the
perturbative FPE, it can be extended to the general FPE without
much difficulty \cite{LH2}.
\bigskip
\section*{Acknowledgments}
This work is supported in part by the
National Science Council (NSC) of the Republic of China under
Grants NSC-99-2112-M-032-002-MY3 and NSC-99-2811-M-032-012.
|
\section{Introduction}
This paper addresses the problems of estimation and inference of parameters
when a group structure among parameters is known. We propose\vadjust{\goodbreak} a new
penalty for
the case where the groups are assumed to gather either nonpositive,
nonnegative or null parameters. All such groups will be referred to as
\textit{sign-coherent}.
As the main motivating example, we consider the linear regression model
\begin{equation}
\label{eqlinearreggroup}
Y = X \bbeta^\star+ \varepsilon
= \sum_{k=1}^K \sum_{j\in\group} X_j \beta_j^\star+ \varepsilon
,
\end{equation}
where $Y$ is a continuous response variable, $X=(X_1,\ldots,X_p)$ is a
vector of $p$
predictor variables, $\bbeta^\star$ is the vector of unknown
parameters and
$\varepsilon$ is a zero-mean Gaussian error variable with variance
$\sigma^2$.
The set of indexes $\{1,\ldots,p\}$ is partitioned into $K$ groups
$\{\group\}_{k=1}^K$ corresponding to predictors and parameters.
We will assume throughout this paper that $\bbeta^\star$ has few nonzero
coefficients, with sparsity and sign patterns governed by the groups~$\group$,
that is, groups being likely to gather either positive, negative or null
parameters.
The estimation and inference of $\bbeta^\star$ is based on training data,
consisting of a vector
$\mathbf{y}=(y_1,\ldots,y_n)^\intercal$ for responses and a
$n\times p$ design matrix $\mathbf{X}$ whose $j$th column contains
$\mathbf{x}_j = (x_j^1,\ldots,x_j^n)^\intercal$, the $n$ observations
for variable $X_j$. For\vspace*{1pt} clarity, we assume that both $\mathbf{y}$
and $\{\mathbf{x}_j\}_{j=1,\ldots,p}$ are centered so as to eliminate the
intercept from fitting criteria.
Penalization methods that build on the $\ell_1$-norm, referred to as
\emph{Lasso} procedures (Least Absolute Shrinkage and Selection
Operator), are
now widely used to tackle simultaneously variable estimation and
selection in
sparse problems. Among these, the group-Lasso, independently proposed by
\citet{1998NIPSGrandvalet} and \citet{1999thesisBakin} and
later developed
by \citet{2006JRSSYuan}, uses the group
structure to define a shrinkage estimator of the form
\begin{equation}
\label{eqgrouplassolinear}
\hat{\bbeta}{}^{\mathrm{group}} = \argmin_{\bbeta\in\mathbb{R}^p}
\Biggl\{
\frac{1}{2} \| \mathbf{y} - \mathbf{X}\bbeta \|^2 +
\lambda\sum_{k=1}^K w_k \|\bbeta_{\mathcal{G}_k} \|
\Biggr\}
,
\end{equation}
where $\mathcal{G}_k$ is the subset of indices defining the $k$th
group of
variables and $\| \cdot \|$ is the Euclidean norm. The tuning
parameter $\lambda\geq 0$ controls the overall amount of penalty and
weights $w_k>0$ adapt the level of penalty within a given
group. Typically, one sets $w_k = \sqrt{p_k}$, where $p_k$ is the
cardinality of~$\group$ in order to adjust
shrinkage according to group sizes. The penalizer in~\eqref{eqgrouplassolinear} is known to induce sparsity at the
group level, setting a whole group of parameters to zero for values of
$\lambda$ which are large enough. Note that when we assign one group
to each predictor, we recover the original Lasso [\citet
{1996JRSSTibshirani}].
The algorithms for finding the group-Lasso estimator have considerably improved
recently.
\citet{2010preprintFoygel} develop a block-wise algorithm, where
each group of
coefficients is updated at a time, using a single line search that
provides the
exact optimal value for one group, considering all other coefficients fixed.
\citet{2008JRSSMeier} depart from linear regression in problem
\eqref{eqgrouplassolinear} by studying group-Lasso penalties for logistic
regression. Their block-coordinate descent method is applicable to
generalized linear models.
Here, we build on the subdifferential calculus approach originally
proposed by
\citet{2000JCGSOsborne} for the Lasso, whose active set algorithm
has been
adapted to the group-Lasso [\citet{2008ICMLRoth}].
Compared to the group-Lasso, this paper deals with a stronger
assumption regarding the group structure. Groups should not only reveal the
sparsity pattern, but they should also be relevant for sign patterns:
all coefficients within a group should be sign-coherent, that is, they should
either be null, nonpositive or nonnegative.
This desideratum arises often when the groups gather redundant or consonant
variables (a usual outcome when groups are defined from clusters of correlated
variables).
To perform this sign-coherent grouped variable selection, we propose a novel
penalty that we call the cooperative-Lasso, in short the \emph{coop-Lasso}.
The coop-Lasso is amenable to the selection of patterns
that cannot be achieved with the group-Lasso. This ability, which can be
observed for finite samples, also leads to consistency results under
the mildest
assumptions.
Indeed, the consistency results for the group-Lasso assume that the
set of nonzero coefficients of $\bbeta^\star$ is
an exact union of groups [\citet{2008JMLRBach}; \citet
{2008EJSNardi}], while
exact support recovery may be achieved with coop-Lasso when some zero
coefficients belong to a group having either positive or negative coefficients.
For example, with groups $\group[1]=\{1,2\}$ and $\group[2]=\{3,4,5\}
$, the
support of $\bbeta^\star=(-1,1,0,1,1)^\intercal$ may be recovered
with the
coop-Lasso, but not with the group-Lasso,
which may then deteriorate the performances of the Lasso [\citet
{2010preprintHuang}].
\citet{2010preprintFriedman} propose to overcome this restriction
by adding an
$\ell_1$ penalty to the objective function in~\eqref
{eqgrouplassolinear}, in
the vein of the hierarchical penalties of \citet{2009ASZhao}.
The new term provides additional flexibility but demands an additional tuning
parameter, while our approach takes a different stance by assuming
sign-coherence, with the benefit of requiring a single tuning parameter.
Section~\ref{secdata} describes two applications where sign-coherence
is a
sensible assumption.
The first one considers ordered categorical data, which are common in
regression and classification. The coop-Lasso can then be used to
induce a monotonic
response to the ordered levels of a covariate, without translating each
level of
the categorical variable into a prescribed quantitative value.
The second application describes the situation where redundancy in measurements
causes sign-coherence to be expected.
Similar behaviors should be observed when features have been grouped by
a clustering
algorithm such as average linkage hierarchical clustering, which are nowadays
routinely used for grouping genes in microarray data analysis
[\citet{1998PNASEisen}; \citet{2006BSPark}; \citet
{2007BMCMa}].\vadjust{\goodbreak}
Finally, in numerous problems of multiple inference, the sign-coherence
assumption is also reasonable: when predicting closely related
responses (e.g.,
regressing male and female life expectancy against economic and social
variables) or when analyzing multilevel data (e.g., predicting academic
achievement against individual factors across schools), the set of coefficients
associated to a predictor (resp., for all response variables or
all data
clusters) forms a group that can often be considered as sign-coherent because
effects can be assumed to be qualitatively similar.
Along these lines, we successfully applied the coop-Lasso penalizer for the
joint inference of several network structures [\citet{2010SCChiquet}].
The rest of the paper is organized as follows: Section~\ref{seccoop}
presents the coop-Lasso penalty, with
the derivation of the optimality conditions which are the basis
for an active set algorithm.
Consistency results and the
associated irrepresentable conditions are given in
Section~\ref{secconsistency}. In
Section~\ref{secmodelselection} we
derive an approximation of the degrees of freedom that
can be used in the Bayesian Information Criterion (BIC) and the Akaike
Information Criterion (AIC)
for model selection. Section~\ref{secnumerical} is dedicated
to simulations assessing the performances of the coop-Lasso in terms of sparsity
pattern recovery, parameters estimation and robustness.
Section~\ref{secdata} considers real data sets, with ordinal and
continuous covariates.
Note that all proofs are postponed until the \hyperref[appm]{Appendix}.
\section{Cooperative-Lasso}\label{seccoop}
\subsection{Definitions and optimality conditions}
\textit{Group-norm and coop-norm.}
We define a group structure by setting a partition of the index set
$\mathcal{I}=\{1,\ldots,p\}$, that is,
\[
\mathcal{I}=\bigcup_{k=1}^K\group \qquad \mbox{with }
\group\cap\group[\ell]=\varnothing
\mbox{ for } k\neq\ell.
\]
Let $\mathbf{v} = (v_1,\ldots,v_p)^\intercal \in \Rset^p$ and
$p_k
$
denote the cardinality of group $k$. We define
$\mathbf{v}_{\group} \in \Rset^{p_k}$ as the vector $(v_j)_{j\in
\group}$.
For the chosen groups $\{\group\}_{k=1}^K$, the group-Lasso norm reads
\begin{equation}
\label{eqgroupnorm}
\| \mathbf{v} \|_{\mathrm{group}} = \sum_{k=1}^K
w_k \| \mathbf{v}_{\group} \|
,
\end{equation}
where $w_k>0$ are fixed parameters enabling to adapt the amount of
penalty for each group. Likewise, the sparse group-Lasso norm
[\citet{2010preprintFriedman}] is defined as a convex combination
of the
group-Lasso and the $\ell_1$ norms:
\begin{equation}
\label{eqsparsegroupnorm}
\| \mathbf{v} \|_{\mathrm{sgl}} = \alpha\| \mathbf{v} \|_{\mathrm{group}}
+ (1-\alpha) \| \mathbf{v} \|_{\mathrm{1}}
,
\end{equation}
where $\alpha$ is meant to be a tuning parameter, but may be fixed to $1/2$
[\citet{2010preprintFriedman}; \citet{Zhou01102010}]. We
will always set it to this
default value in what follows.
Let $\mathbf{v}^+ = (v^+_1,\ldots,v^+_p)^\intercal$ and
$\mathbf{v}^- = (v^-_1,\ldots,v^-_p)^\intercal $ be the componentwise
positive and negative part of $\mathbf{v}$, that is, $ v^+_j =
\max(0,v_j)$ and $v^-_j = \max(0,-v_j)$, respectively. We call
\emph{coop-norm} of $\mathbf{v}$ the sum of group-norms on
$\mathbf{v}^+$ and $\mathbf{v}^-$,
\[
\|\mathbf{v} \|_{\mathrm{coop}} =
\|\mathbf{v}^+ \|_{\mathrm{group}} +
\|\mathbf{v}^- \|_{\mathrm{group}}=
\sum_{k=1}^K w_k ( \|\mathbf{v}_{\group}^+ \| +
\|\mathbf{v}_{\group}^- \| )
,
\]
which is clearly a norm on $\mathbb{R}^p$.
The coop-Lasso
estimate of $\bbeta^\star$ as defined in \eqref{eqlinearreggroup}
is
\begin{equation}
\label{eqcooplassolinear}
\hatbbetacoop = \argmin_{\bbeta\in\mathbb{R}^p}
L(\bbeta) \qquad\mbox{with } L(\bbeta) =
\frac{1}{2} \|\mathbf{y}- \mathbf{X}\bbeta \|^
+ \lambd
\| \bbeta\|_{\mathrm{coop}}
,
\end{equation}
where $\lambda \geq 0$ is a tuning parameter common to all
groups. Appropriate choices for $\lambda$
will be discussed in Sections \ref{secmodelselection} and \ref{secconsistency}
dealing with model selection and consistency, respectively.
Illustrations of the group, sparse group and coop norms are given in
Figure~\ref{figcoop-norm} for a vector
$\bbeta= (\beta_{1},\beta_{2},\beta_{3},\beta
_{4})^\intercal$
with two groups $\group[1] = \{1,2\}$ and $\group[2] = \{3,4\}$.
We represent several views of the unit ball for each of these norms.
For the coop-norm, this ball represents the set of feasible
solutions for an optimization problem equivalent to~\eqref{eqcooplassolinear},
where the sum of squared residuals is minimized under unitary
constraints on
$\| \bbeta\|_{\mathrm{coop}}$. The same interpretation holds for the
group and
sparse group norms, provided the sum of squared residuals is minimized
under unitary constraints on $\| \bbeta\|_{\mathrm{group}}$ and $\|
\bbeta\|_{\mathrm{sgl}}$, respectively.
\begin{figure}[t!]
\includegraphics{520f01.eps}
\caption{Feasible sets for the coop-Lasso, group-Lasso and sparse
group-Lasso penalties.
First column: cuts through $(\beta_{1},\beta_{2},\beta_{3})$ at
$\beta_{4}=0$ and $\beta_{4}=0.3$:
$(\beta_{1},\beta_{2})$ span the horizontal plane and $\beta_{3}$ is
on the vertical axis; second and third columns: cuts through
$(\beta_{1},\beta_{3})$ at various values of $(\beta_{2},\beta_{4})$;
last column: cuts through $(\beta_{1},\beta_{2})$ at various values
of $(\beta_{3},\beta_{4})$.}
\label{figcoop-norm}
\end{figure}
These plots provide some insight into the sparsity pattern that
originates from
the penalties, since sparsity is related to the singularities of the
boundary of
the feasible set.
First, consider the group-Lasso: the first row illustrates that when
$\beta_{4}$
is null its group companion $\beta_{3}$ may also be exactly zero
(corners on the
boundary at $\beta_{3}=0$), while the second row shows that this event is
improbable when $\beta_{4}$ differs from zero (smooth boundary at
$\beta_{3}=0$).
The second and third columns display the same type of relationships within
$\group[1]$ between $\beta_{2}$ and $\beta_{1}$, which are expected
due to the
symmetries of the unit ball.
The last column displays $\ell_2$ balls, which characterize the within-groups
feasibility subsets, showing that once a group is activated, all its members
will be nonzero.
Now, consider the sparse group-norm: the combination of the group and Lasso
penalties has uniformly shrunk the feasible set toward the Lasso $\ell
_1$ unit
ball, thus creating new edges that provide a chance to zero any
parameter in any
situation, with an elastic-net-like penalty [\citet{2005JRSSZou}]
within and
between groups. The comparison of the last two columns illustrates that the
differentiation between the within-group and between group penalties is less
marked than for the group-Lasso.
Finally, consider the coop-norm: compared to the group-norm,
there are also additional discontinuities resulting in new edges on the
3-D plots.
While the sparse group-Lasso edges where created by a uniform shrinking
toward the $\ell_1$ unit ball, the coop-Lasso new edges result from slicing
the group-Lasso unit ball, depriving sign-incoherent orthants from some
of the
group-Lasso feasible solutions ($\|\bbeta\|_{\mathrm{coop}} > \| \bbeta
\|_{\mathrm{group}}$ in these regions).
Note that, in general, there are less new edges than with the sparse
group-Lasso, since the new opportunities to zero some coefficients are limited
to the case where the group-Lasso would have allowed a solution with opposite
signs within a group.
The crucial difference with the group and sparse group-Lasso is the
loss of the axial
symmetry when some variables are nonzero:
decoupling the positive and negative parts of the regression
coefficients favors
solutions where signs match within a group.
Slicing of the unit group-norm ball does not affect the positive
and negative orthants, but large areas corresponding to sign mismatches have
been peeled off, as best seen on the last column, which also
illustrates the
strong differentiation between within-group and between-group penalties.
Before stating the optimality conditions for
problem~\eqref{eqcooplassolinear}, we introduce some notation related
to the
sparsity pattern of parameters, which will be required to express the necessary
and sufficient condition for optimality.
First, we recall that the unknown vector of parameters $\bbeta^\star$
is typically sparse; its support is denoted $\supp=\{j, \bbeta
_j^\star\neq0\}$
and $\supp^c=\{j, \bbeta_j^\star= 0\}$ is the complementary set of
true zeros.
Once the problem has been supplied with a group structure, we define
$\supp_k =
\supp\cap\group$ and $\supp_k^c = \supp^c \cap\group$ as the
sets of
relevant, respectively irrelevant, predictors within group $k$, for all
$k=1,\ldots,K$. Similar notation $\supp(\bbeta)$, $\supp_k(\bbeta)$ and
$\supp_k^c(\bbeta)$ is defined for an arbitrary vector $\bbeta\in
\Rset^p$.
Furthermore, for clarity and brevity, we introduce the functions
$\{\bvarphi_j\}_{j=1}^{p}$,
which return the componentwise positive or negative part of a vector according
to the sign of its $j$th elemen
, that is,
$\forall k \in\{1,\ldots,K\} , \forall j\in\group , \forall
\mathbf{v}\in\Rset^{p_k},$
\begin{equation}
\label{eqphij}
\bvarphi_j(\mathbf{v}) = (\sign(v_j)\mathbf{v})^+ =
\cases{\displaystyle
\mathbf{0} ,&\quad if $v_j = 0$,\cr\displaystyle
\mathbf{v}^+ ,&\quad if $v_j > 0$,\cr\displaystyle
\mathbf{v}^- ,&\quad if $v_j < 0$.
}
\end{equation}
\textit{Optimality conditions.}
The objective function $L$ in \eqref{eqcooplassolinear} is
continuous and coercive, thus problem~\eqref{eqcooplassolinear}
admits at least one minimum.
If $\mathbf{X}$ has rank~$p$, then the minimum is unique since~$L$ is strictly
convex.
Furthermore,~$L$ is smooth, except at some locations
with zero coefficients, due to the singularities of the coop-norm.
Since $L$ is convex, a necessary and sufficient condition for the
optimality of $\bbeta$ is that the null vector $\mathbf{0}$ belongs to
the subdifferential of $L$ whose expression is provided in the
following lemma.
\begin{lemma}\label{lemsubdifferential}
For all $\bbeta\in\Rset^p$, the subdifferential of the objective
function of
problem~\eqref{eqcooplassolinear} is
\begin{equation}
\partial_{\bbeta} L(\bbeta) =
\{ \mathbf{v} \in\Rset^p \dvtx
\mathbf{v} = \mathbf{X}^\intercal
(\mathbf{X}\bbeta-\mathbf{y}) + \lambda
\btheta \},
\end{equation}
where $\btheta\in\Rset^p$ is any vector belonging to the
subdifferential of
the coop-norm, that is,
\begin{subequation}
\label{eqsubgradientcompact}
\begin{eqnarray}\label{eqsubgradientcompactb}
\forall k &\in&\{1,\ldots,K\} , \forall j\in\supp_k(\bbeta) \qquad \theta_j = \frac{w_k\beta_j} { \|
\bvarphi_j(\bbeta_{\group}) \|} ,
\\\label{eqsubgradientcompacta}
\forall k &\in&\{1,\ldots,K\} , \forall j\in\supp_k^c(\bbeta) \qquad \| \bvarphi_j (\btheta_{\group} ) \|
\leq w_k .
\end{eqnarray}
\end{subequation}
\end{lemma}
The following optimality conditions, which result directly from
Lemma~\ref{lemsubdifferential}, are an essential building block of the
algorithm we propose to compute the coop-Lasso estimate. They also
provide an
important basis for showing the consistency results.
\begin{theorem}\label{thmoptimality}
Problem \eqref{eqcooplassolinear} admits at least one solution, which is
unique if $\mathbf{X}$ has rank $p$.
All critical points $\bbeta$ of the objective function $L$ verifying the
following conditions are global minima:
\begin{subequation}
\label{eqoptimalityall}
\begin{eqnarray}
\label{eqoptimality}
\forall k &\in&\{1,\ldots,K\} , \forall j\in\supp_k(\bbeta) \qquad \mathbf{x}^\intercal_j(\mathbf{X}\bbeta-\mathbf{y}) +
\frac{\lambda w_k\beta_j}{\|\bvarphi_j(\bbeta_{\group})\|} = 0
, \\
\label{eqoptimalityzero}
\forall k &\in&\{1,\ldots,K\} , \forall j\in\supp_k^c(\bbeta) \qquad \bigl\|
\bvarphi_j \bigl((\mathbf{X}_{\centerdot\group})^{\intercal
}(\mathbf{X}\bbeta-\mathbf{y}) \bigr)
\bigr\| \leq\lambda w_k
,
\end{eqnarray}
\end{subequation}
where $\mathbf{X}_{\centerdot\group}$ is the submatrix of $\mathbf
{X}$ with
all rows and columns indexed by $\group $.
\end{theorem}
Note here an important distinction compared to the group-Lasso, where
the optimality
conditions are expressed solely according to the groups $\group$
[see, e.g., \citet{2008ICMLRoth}].
Hence, while the sparsity pattern of the solution is strongly
constrained by the
predefined group structure in the group-Lasso, deviations from this structure
are possible for the coop-Lasso.
The asymptotic analysis of Section~\ref{secconsistency} confirms that exact
support recovery is possible even when the support of $\bbeta^\star$
cannot be
expressed as a simple union of groups, provided the groups intersecting
the true
support are sign-coherent.
\subsection{Algorithm}
The efficient approaches developed for the Lasso take advantage of the sparsity
of the solution by solving a series of small linear systems, whose
sizes are
incrementally increased/decreased [\citet{2000JCGSOsborne}]. This
approach was
pursued for the group-Lasso [\citet{2008ICMLRoth}] and we
proposed an
algorithm in the same vein for the coop-Lasso in the framework of
multiple network inference [\citet{2010SCChiquet}].
We provide here a more detailed description of the latter in the
specific context
of linear regression.
The algorithm starts
from a sparse initial guess, say, $\bbeta=0$, and iterates two steps:\vadjust{\goodbreak}
\renewcommand\thelonglist{\arabic{longlist}}
\renewcommand\labellonglist{\arabic{longlist}.}
\begin{longlist}[2.]
\item\label{itemalgostep1} The first step solves problem~\eqref
{eqcooplassolinear} with respect to
$\bbeta_{ \mathcal{A}}$, the subset of ``active'' variables, currently
identified as being nonzero.
At this stage the current feasible set is restricted to the orthants
where the
gradient of the coop-norm has no discontinuities: the optimization
problem is
thus smooth.
One or more variables may then be declared inactive if the current optimal~$\bbeta_{ \mathcal{A}}$ reaches the boundary of the current feasible set.
\item\label{itemalgostep2} The second step assesses the completeness
of the set $\mathcal{A}$,
by checking the optimality conditions with respect to inactive
variables. We add a group
that violates these conditions.
In our implementation, we pick the one that most violates the optimality
condition, since this strategy has been observed to require few changes
in the active set.
When no such violation exists, the current solution is optimal.
\end{longlist}
\begin{algorithm}[t]
\hspace*{22pt}\begin{minipage}{327pt}
\nlset{{\normalsize Init.}} Start from a feasible $\bbeta\leftarrow\bbeta^0$
\begin{eqnarray*}
\mathcal{A}_{+} &\leftarrow&\{j\in\group\dvtx \|\bbeta_{\group}^+\| >
0 , k=1,\ldots,K\}, \\
\mathcal{A}_{-} &\leftarrow&\{j\in\group\dvtx \|\bbeta_{\group}^-\| >
0 , k=1,\ldots,K\}.
\end{eqnarray*}
\nlset{{\normalsize Step \ref{itemalgostep1}}}On $\mathcal{A} \leftarrow\mathcal
{A}_{+} \cup\mathcal{A}_{-}$, find
a solution to the smooth problem
\begin{eqnarray}
\bbeta_{ \mathcal{A}} \leftarrow
\argmin_{\mathbf{v}\in\mathbb{R}^{|\mathcal{A}|}} \frac
{1}{2} \|\mathbf{y}-\mathbf{X}_{\centerdot\mathcal{A}} \mathbf
{v} \|^2
+ \lambda \| \mathbf{v} \|_{\mathrm{coop}}\nonumber\\
\eqntext{\mbox{s.t. }
\cases{\displaystyle
v_j \geq0 ,&\quad if $j\in\mathcal{A}_{+} \cap\mathcal{A}_{-}^c
$,\cr\displaystyle
v_j \leq0 ,&\quad if $ j\in\mathcal{A}_{-} \cap\mathcal{A}_{+}^c
$,
}}
\end{eqnarray}
where $\mathcal{A}_{-}^c$ and $\mathcal{A}_{+}^c$ are the
complementary sets
of $\mathcal{A}_{-}$ and $\mathcal{A}_{+}$, respectively.
Identify groups inactivated during optimization
\begin{eqnarray*}
\mathcal{A}_{+} &\leftarrow&\mathcal{A}_{+}\bigm\backslash\Bigl\{
j\in\group\subseteq\mathcal{A}_{+}\dvtx \|\bbeta_{\group}^+\| = 0
\\
&&\hphantom{\mathcal{A}_{+}\setminus\Bigl\{} \mbox{and } \min_{\mathbf{v} \in\partial_{\bbeta_{\group}}
L(\bbeta)} \|\mathbf{v^-} \|=0 , k=1,\ldots,K\Bigr\} ,\\
\mathcal{A}_{-} &\leftarrow&\mathcal{A}_{-}\bigm\backslash\Bigl\{
j\in\group\subseteq\mathcal{A}_{-}\dvtx \|\bbeta_{\group}^-\| = 0
\\
&&\hphantom{\mathcal{A}_{-}\bigm\backslash\Bigl\{}\mbox{and } \min_{\mathbf{v} \in\partial_{\bbeta_{\group}}
L(\bbeta)} \|\mathbf{v^+} \|=0 , k=1,\ldots,K\Bigr\}.
\end{eqnarray*}
\nlset{{\normalsize Step \ref{itemalgostep2}}}
Identify the greatest violation of optimality conditions:
\begin{eqnarray*}
g^k_+ &\leftarrow&\min_{\mathbf{v} \in\partial
_{\bbeta_{\group}} L(\bbeta)} \|\mathbf{v^+} \|
, \qquad
q \leftarrow\argmax_{k} g^k_+ , \\
g^k_- &\leftarrow&\min_{\mathbf{v} \in\partial
_{\bbeta_{\group}} L(\bbeta)} \|\mathbf{v^-} \|
, \qquad
r \leftarrow\argmax_{k} g^k_-
\end{eqnarray*}
\eIf{$\max(g^q_+,g^r_-)=0$}{
Stop and return $\bbeta$, which is
optimal
}{
\lIf{$g^q_+ > g^r_-$}{ $\mathcal{A}_{-} \leftarrow\mathcal{A}_{-}
\cup
\group[q]$ }
\lElse{ $\mathcal{A}_{+} \leftarrow\mathcal{A}_{+} \cup\group[r]$ }\\
Repeat Steps 1 and 2 until convergence }
\vspace*{1pt}
\end{minipage}
\caption{Coop-Lasso fitting algorithm}
\label{algoactiveconstraint}
\end{algorithm}
These two steps outline the algorithm, which is detailed in more
technical terms in Algorithm~\ref{algoactiveconstraint}. The
principle is readily applied to any generalized linear model by simply
defining the appropriate objective function $L$. In our current
implementation (a pre-release of our \texttt{R}-package \texttt{scoop}
is available at \url{http://stat.genopole.cnrs.fr/logiciels/scoop})
the linear and logistic regression models are implemented using either
Broyden--Fletcher--Goldfarb--Shanno (BFGS) quasi-Newton updates with
box constraints, or proximal methods [\citet{2009SIAMBeck}] to solve
the smooth optimization problem in Step~\ref{itemalgostep1}.
Finally, note that to compute a series of solutions along the
regularization path for
problem~\eqref{eqcooplassolinear}, we simply choose a series of penalties
$\lambda^1=\lambda_{\mathrm{max}} >\cdots > \lambda^l>\cdots > \lambda
^L =
\lambda_{\mathrm{min}} \geq0$ such that $\hatbbetacoop(\lambda
_{\mathrm{max}})=\mathbf{0}$, that is,
\[
\lambda_{\mathrm{max}} = \max_{k\in\{1,\ldots,K\}} \max_{j\in
\group} \frac{1}{w_k}
\|\bvarphi_j ((\mathbf{X}_{\centerdot\group})^\intercal
\mathbf{y} ) \|
.
\]
We then use the usual warm start strategy, where the feasible initial
guess for
$\hatbbetacoop(\lambda^{l})$, the coop-Lasso estimate with penalty parameter
$\lambda^l$, is initialized with $\hatbbetacoop(\lambda^{l-1})$.
\subsection{Orthonormal design case}
The orthonormal design case, where\break $\mathbf{X}^\intercal\mathbf{X} =
\mathbf{I}_p$, has been providing useful insights for penalization techniques
regarding the effects of shrinkage.
Indeed, in this particular case, most usual shrinkage estimators can be
expressed in closed-form as functions of the ordinary least squares (OLS)
estimate.
These expressions pave the way for the derivation of approximations of the
degrees of freedom
[\citet{1996JRSSTibshirani}; \citet{2006JRSSYuan} and
Section~\ref{secmodelselection}], which
may be convenient for model selection in the absence of exact formulae.
In the orthonormal setting, for any $\beta_j$, we have
$\mathbf{x}_j^\intercal(\mathbf{X}\bbeta-\mathbf{y}) = \beta_j -
\hatbeta_j^{\mathrm{ols}}$.
The optimality conditions~\eqref{eqoptimality} and~\eqref{eqoptimalityzero}
can then be written as
\begin{equation}
\label{eqcooportho}
\forall k \in\{1,\ldots,K\} , \forall j\in\group \qquad \hatbeta_j^{\mathrm{coop}} = \biggl(
1-\frac{\lambda w_k}{ \|\bvarphi_j(\hatbbetaols_{\group}) \|}
\biggr)^{ +}
\hatbeta_j^{\mathrm{ols}}
.
\end{equation}
For reference, we recall the solution to the group-Lasso [\citet
{2006JRSSYuan}] in the
same condition
\begin{equation}
\label{eqgrouportho}
\forall k \in\{1,\ldots,K\} , \forall j\in\group \qquad
\hatbeta_j^{\mathrm{group}} = \biggl (
1-\frac{\lambda w_k}{ \| \hatbbetaols_{\group} \|}
\biggr)^{ +}
\hatbeta_j^{\mathrm{ols}}
,
\end{equation}
while the Lasso solution [\citet{1996JRSSTibshirani}] is
\begin{equation}
\label{eqlassoortho}
\forall j\in\{1,\ldots,p\} \qquad
\hatbeta_j^{\mathrm{lasso}} = \biggl(
1-\frac{\lambda}{ | \hatbeta_j^{\mathrm{ols}} |}
\biggr)^{ +}
\hatbeta_j^{\mathrm{ols}}
.
\end{equation}
Equations~\eqref{eqcooportho}--\eqref{eqlassoortho} reveal strong
commonalities.
First, the coefficients of these shrinkage estimators are of the sign
of the OLS estimates.
Second, the norm used in the penalty defines a region where small OLS
coefficients are shrunk to zero, while large ones are shrunk inversely
proportional to this norm.
Finally, by grouping the terms corresponding to one
group in equations~\eqref{eqcooportho}--\eqref{eqgrouportho}, a uniform
translation effect, analogous to the one observed for the Lasso, comes into
view:
\begin{eqnarray}
\label{eqcoopnormortho}
&\displaystyle \forall k \in \{1,\ldots,K\} , \forall j\in\group \qquad
\|\bvarphi_j(\hatbbetacoop_{\group}) \| = \bigl(
\| \bvarphi_j(\hatbbetaols_{\group}) \| - \lambda w_k
\bigr)^{ +}
,& \nonumber\\
&\displaystyle \forall k \in \{1,\ldots,K\} \qquad
\| \hatbbetagroup_{\group} \| = (
\| \hatbbetaols_{\group} \|- \lambda w_k
)^{ +}
,&\\
&\displaystyle \forall j \in \{1,\ldots,p\} \qquad
| \hatbeta_j^{\mathrm{lasso}} | = (
| \hatbeta_j^{\mathrm{ols}} | - \lambda w_k
)^{ +}
.&\nonumber
\end{eqnarray}
The group-Lasso \eqref{eqgrouportho} differs primarily
from the Lasso \eqref{eqlassoortho} owing to the common penalty
${\lambda
w_k}/{\| \hatbbetaols_{\group}\|}$ for all the coefficients belonging
to group
$k$. The magnitude of shrinkage is determined by all within-group OLS
coefficients,
and is thus radically different from a ridge regression penalty in this regard.
For the coop-Lasso estimator \eqref{eqcooportho}, two penalties
possibly apply
to group $k$, for the positive and the negative OLS coefficients, respectively.
If all within-group OLS coefficients are of the same sign, coop-Lasso
is identical
to group-Lasso; if some signs disagree, the magnitude of the penalty
only depends on the within-group OLS coefficients with an identical
sign. In the
extreme case where exactly one OLS coefficient is positive/negative, the
coop-penalty is identical to a Lasso penalty on this coefficient.
Note that such a simple analytical formulation is not available for the sparse
group-Lasso estimate $\hatbbeta{}^{\mathrm{sgl}}$, but an expression can
be obtained
by chaining two simple shrinkage operations.
Introducing an intermediate solution $\tilde{\bbeta}{}^\mathrm{sgl}$, we have,
$\forall k \in\{1,\ldots,K\}$ and $\forall j\in\group ,$
\begin{equation}
\label{eqsparsegrouportho}\qquad
\hatbeta_j^{\mathrm{sgl}} = \biggl(
1-\frac{\lambda(1-\alpha) w_k}{ \|\tilde{\bbeta}{}^\mathrm
{sgl}_{\group} \|}
\biggr)^{ +}
\tilde{\beta}{}^\mathrm{sgl}_{j}
\qquad \mbox{where }
\tilde{\beta}{}^\mathrm{sgl}_{j} = \biggl(
1-\frac{\lambda\alpha}{ | \hatbeta_j^{\mathrm{ols}} |}
\biggr)^{ +}
\hatbeta_j^{\mathrm{ols}}
.
\end{equation}
The intermediate solution $\tilde{\bbeta}{}^\mathrm{sgl}$ is the Lasso estimator
with penalty parameter~$\lambda\alpha$, which acts as the OLS
estimate for a
group-Lasso of parameter $\lambda(1-\alpha)$.
Figure~\ref{figorthonormal} provides a visual representation of equations
\eqref{eqcooportho}--\eqref{eqlassoortho} and \eqref{eqsparsegrouportho}
for a group with two components, say, $\group=\{1,2\}$.
\begin{figure}
\includegraphics{520f02.eps}
\caption{Lasso, group, sparse group and coop Lasso
coefficient estimates, for a group with 2 elements $\group=\{1,2\}$,
as a
function of the OLS coefficients. The colors emphasize the positive and
negative quadrants of the
$(\hatbeta{}^{\mathrm{ols}}_1,\hatbeta{}^{\mathrm{ols}}_2)$ plane, with red and
blue,
respectively.}\label{figorthonormal}\vspace*{-3pt}
\end{figure}
We plot $\hatbeta_1^{\mathrm{lasso}},
\hatbeta_1^{\mathrm{group}}, \hatbeta_1^{\mathrm{sgl}}$ and $\hatbeta
_1^{\mathrm{coop}}$ as functions of
$(\hatbeta_1^{\mathrm{ols}},\hatbeta_2^{\mathrm{ols}})$.
Top-left, the Lasso translates the $\hatbeta_1^{\mathrm{ols}}$
coefficient toward zero, eventually truncating them at zero, regardless of~$\hatbeta_2^{\mathrm{ols}}$: there is no interaction between coefficients.
The group-Lasso, top-right, has a nonlinear shrinking behavior (quite
different from the Lasso or ridge penalties in this respect) and sets
$\hatbeta_1^{\mathrm{group}}$ to zero within a Euclidean ball centered
at zero.
The sparse group-Lasso, bottom-left, is a hybrid of Lasso and
group-Lasso, whose shrinking behavior lies between its two ancestors.
Bottom-right, the coop-Lasso appears as another form of cross-breed, identical
to the group-Lasso in the positive and negative quadrants, and
identical to the
Lasso when the signs of the OLS coefficients mismatch.
For groups with more than two components, intermediate solutions would be
possible.
This behavior is shown to allow for some flexibility
with respect to the predefined group structure in the following consistency
analysis.
\section{Consistency}\label{secconsistency}
Beyond its sanity-check value, a consistency analysis brings along an
appreciation of the strengths and limitations of an estimation scheme.
Here we concentrate on the estimation of the support of the parameter
vector, that is, the position of its zero entries.
Our proof technique is drawn from the previous works on the Lasso
[\citet{2007JRSSYuan}] and the group-Lasso
[\citet{2008JMLRBach}].\vadjust{\goodbreak}
In this type of analysis, some assumptions on the joint distribution
of $(X,Y)$ are required to guarantee the convergence of empirical covariances.
For the sake of simplicity and coherence, we keep assuming that data are
centered so that we have zero mean random variables and
$\bPsi= \mathbb{E} [XX^\intercal ]$ is the covariance
matrix of $X$:
\begin{longlist}[(A2)]
\item[(A1)] $X$ and $Y$ have finite 4th order
moments $\mathbb{E} [\|X\|^4 ] < \infty$, $\mathbb
{E} [Y^4 ] < \infty$.
\item[(A2)] The covariance matrix $\bPsi= \mathbb{E}
[XX^\intercal ] \in\Rset^{p \times p}$ is invertible.
\end{longlist}
In addition to these standard technical assumptions, we need a more specific
one, substantially avoiding situations where the coop-Lasso will almost never
recover the true support:
\begin{longlist}[(A3)]
\item[(A3)]
All sign-incoherent groups are included in the true
support:
$\forall k\in\{1,\ldots,K\}$, if
$\|(\bbeta_{\group}^\star)^+\| >0$ and
$\|(\bbeta_{\group}^\star)^-\| >0$, then $\forall j \in\group$,
$\bbeta^\star_j \neq0$.
\end{longlist}
Note that this latter assumption is less stringent than the one
required for the
group-Lasso since it does not require that each group of variables
should either
be included in or excluded from the support.
For the coop-Lasso, sign-coherent groups may intersect the support.
The spurious relationships that may arise from confounding variables are
controlled by the so-called strong irrepresentable condition, which guarantees
support recovery for the Lasso [\citet{2007JRSSYuan}] and the group-Lasso
[\citet{2008JMLRBach}].
We now introduce suitable variants of these conditions for the coop-Lasso.
They result in two assumptions: a general one, on the
magnitude of correlations between relevant and irrelevant variables,
and a~more
specific one for groups which intersect the support, on the sign of
correlations.
These conditions will be expressed in a compact vectorial form using the
diagonal weighting matrix $\mathbf{D}(\bbeta)$ such that,
\begin{equation}\label{eqdefD}
\forall k \in\{1,\ldots,K\}, \forall j\in\supp_k(\bbeta) \qquad
(\mathbf{D}(\bbeta) )_{jj} = w_k \|\bvarphi_j(\bbeta
_{\group})\|^{-1}
.
\end{equation}\vspace*{-\baselineskip}
\begin{longlist}[(A4)]
\item[(A4)] For every group $\group$ including at least one null
coefficient (i.e., such that $\beta^\star_{j} = 0$ for some $j \in\group$ or, equivalently,
$\supp^c_k \neq\varnothing$), there exists $\eta>0$ such that
\begin{equation}
\frac{1}{w_k} \max( \| (\bPsi_{\supp_k^c\supp} \bPsi_{\supp\supp}^{-1}
\mathbf{D}(\bbeta^\star_{\supp})\bbeta^\star_{\supp})^+ \|,
\| (\bPsi_{\supp_k^c \supp} \bPsi_{\supp\supp}^{-1}
\mathbf{D}(\bbeta^\star_{\supp})\bbeta^\star_{\supp})^-
\| ) \leq1-\eta
, \label{thICnoPnoN}\hspace*{-35pt}
\end{equation}
where $\bPsi_{\cal{ST}}$ is the submatrix of $\bPsi$ with lines and columns
respectively indexed by
$\cal{S}$ and $\cal{T}$
\end{longlist}
\begin{longlist}[(A5)]
\item[(A5)] For every group $\group$ intersecting the support and including
either positive or negative coefficients, letting $\nu_k$ be the sign
of these
coefficients [$\nu_k = 1$ if $\|(\bbeta_{\group}^\star)^+\| >0$ and
$\nu_k=-1$
if $\|(\bbeta_{\group}^\star)^-\| >0$], the following inequalities should
hold:
\begin{equation}
\nu_k \bPsi_{\supp^c_k \supp} \bPsi_{\supp\supp}^{-1}
\mathbf{D}(\bbeta^\star_{\supp})\bbeta^\star_{\supp}
\preceq\mathbf{0}
,
\label{thICPorN}
\end{equation}
where $\preceq$ denotes componentwise inequality.\vadjust{\goodbreak}
\end{longlist}
Note that the irrepresentable condition for the group-Lasso only considers
correlations between groups included and excluded from the support.
It is otherwise similar to \eqref{thICnoPnoN}, except that the
elements of the
weighting matrix~$\mathbf D$ are $w_k \|\bbeta_{\group}\|^{-1}$ and
that the
$\ell_2$ norm replaces $ \max( \| (\cdot)^+ \|, \| (\cdot)^-\| ) $.
We now have all the components for stating the coop-Lasso consistency theorem,
which will consider the following normalized (equivalent) form of the
optimization problem \eqref{eqcooplassolinear} to allow a direct comparison
with the known similar results previously stated for the Lasso and group-Lasso
[\citet{2007JRSSYuan}; \citet{2008JMLRBach}]:
\begin{equation}
\hatbbetacoop_n = \argmin_{\bbeta\in\mathbb{R}^p}
\frac{1}{2n} \|\mathbf{y}- \mathbf{X}\bbeta \|^2
+ \lambda_n
\| \bbeta\|_{\mathrm{coop}}
,
\label{eqcooplassonormalized}
\end{equation}
where $\lambda_n = \lambda/n$.
\begin{theorem} \label{thmsupportconsistency}
If assumptions \textup{(A1)--(A5)} are satisfied, the coop-Lasso estimator is asymptotically
unbiased and has the property of exact support recovery:
\begin{equation}
\hatbbetacoop_n \inprob\bbeta^\star\quad\mbox{and} \quad
\prob \bigl(\supp(\hatbbetacoop_n) = \supp \bigr) \rightarrow1
, \label{eqconsistency}
\end{equation}
for every sequence $\lambda_n$ such that
$\lambda_n = \lambda_0
n^{- \gamma}, \gamma\in(0,1/2) $.
\end{theorem}
Compared to the group-Lasso, the consistency of support recovery for the
coop-Lasso differs primarily regarding possible intersection (besides inclusion
and exclusion) between groups and support.
This additional flexibility applies to every sign-coherent group.
Even if the support is the union of groups, when all groups are sign-coherent,
the coop-Lasso has still an edge on group-Lasso since the irrepresentable
condition \eqref{thICnoPnoN} is weaker.
Indeed, the norm in \eqref{thICnoPnoN} is dominated by the $\ell_2$
norm used
for the group-Lasso.
The next paragraph illustrates that this difference can have remarkable
outcomes.
Finally, when the support is the union of groups comprising
sign-incoherent ones,
there is no systematic advantage in favor of one or the other method.
While the
norm used by the coop-Lasso is dominated by the norm used by the group-Lasso,
the weighting matrix $\mathbf{D}$ has smaller entries for the latter.
\textit{Illustration.}
We generate data from the regression model \eqref{eqlinearreggroup}, with
$\bbeta^\star= (1,1,-1,-1,0,0,0,0)$, equipped with the group
structure $\{\group\}_{k=1}^4 = \{\{1,2\},\allowbreak \{3,4\}, \{5,6\},\{
7,8\}\}$.
The vector $X$ is generated as a centered Gaussian random vector whose
covariance matrix $\bPsi$ is chosen so that the irrepresentable
conditions hold
for the coop-Lasso, but not for the group-Lasso, which, we recall, are more
demanding for the current situation, with sign-coherent groups.
The random error $\varepsilon$ follows a centered Gaussian
distribution with
standard deviation $\sigma=0.1$, inducing a very high signal to noise
ratio ($R^2=0.99$ on average), so that asymptotics provide a realistic
view of
the finite sample situation.
We generated 1000 samples of size $n=20$ from the described model, and
computed the
corresponding 1000 regularization paths for the group-Lasso, sparse group-Lasso
and coop-Lasso.
Figure~\ref{figconsistencyillustration} reports the 50\% coverage intervals
(lower and upper quartiles) along the regularization paths.
In this setup, the sparse group-Lasso behaves as the group-Lasso,
leading to
nearly identical graphs.
\begin{figure}
\includegraphics{520f03.eps}
\caption{50\% coverage intervals for the group (left), sparse group
(center) and (right) Lasso estimated coefficients along regularization
paths: coefficients from the support of $\bbeta^\star$ are marked by
colored horizontal stripes and the other ones by gray vertical stripes.}
\label{figconsistencyillustration}
\end{figure}
Estimation is difficult in this small sample problem
($n=20, p=8$), and the two
versions of the group-Lasso, which first select the wrong covariates, never
reach the situation where they would have a decisive advantage upon
OLS, while
the coop-Lasso immediately selects
the right covariates, whose coefficients steadily dominate the
irrelevant ones.
Model selection is also difficult, and the BIC criteria provided in
Section~\ref{secmodelselection} select often the OLS model (in about
$10\%$
and $50\%$ of cases for the coop-Lasso and the group-Lasso, respectively).
The average root mean square error on
parameters is of order $10^{-1}$ for all methods, with a slight edge
for the coop-Lasso.
The sign error is much more contrasted: $31\%$ for the coop-Lasso
\textit{vs.}
$46\%$ for the group-Lasso, not far better than the $50\%$ of OLS.
\section{Model selection}\label{secmodelselection}
Model selection amounts here to choosing the penalization parameter
$\lambda$,
which restricts the size of the
estimate $\hatbbeta(\lambda)$. Trial values
$\{\lambda_\mathrm{min},\ldots,\lambda_\mathrm{max}\}$ define the
set of models
we have to choose from along the regularization path.
The process aims at picking the model with minimum prediction error, or the
one closest to the model from which data have been generated,
assuming the model is correct, that is, equation~\eqref{eqlinearreggroup}
holds.
Here ``closest'' is typically measured by a distance between $\hatbbeta
$ and
$\bbeta^\star$, either based on the value of the coefficients or on their
support (true model selection), and sometimes also
on the sign correctness of each nonzero entry.
Among the prerequisite for the selection process to be valid,
the previous consistency analysis comes
up with suitable orders of magnitude for the penalty parameter $\lambda
.
However, it does not provide a proper value to be plugged in
\eqref{eqcooplassolinear}
and the practice is to
use data driven approaches for selecting an appropriate penalty parameter.
Cross-validation is a recommended option [\citet{Hesterberg08}]
when looking for
the model minimizing the prediction error, but it is slow and not well
suited to
select the model closest to the true one.
Analytical criteria provide a faster way to perform model selection and,
though the information criteria AIC and BIC rely on asymptotic
derivations, they
often offer good practical performances.
The BIC and AIC criteria for the Lasso [\citet{2007ASZou}] and group-Lasso
[\citet{2006JRSSYuan}] have been defined through the effective
degrees of
freedom:
\begin{eqnarray}
\label{eqAICcoop}
\mathrm{AIC}(\lambda) & =& \frac{\|\mathbf{y} -
\hat{\mathbf{y}}(\lambda)\|^2}{\sigma^2}+ 2 \operatorname{df}(\lambda)
, \\
\label{eqBICcoop}
\mathrm{BIC}(\lambda) & =& \frac{\|\mathbf{y} -
\hat{\mathbf{y}}(\lambda)\|^2}{\sigma^2}+\log(n)
\operatorname{df}(\lambda)
,
\end{eqnarray}
where $\hat{\mathbf{y}}(\lambda)=\mathbf{X}\hatbbeta(\lambda)$ is
the vector of
predicted values for \eqref{eqcooplassolinear} with penalty parameter
$\lambda$,
$\sigma^2$ is the variance of the zero-mean Gaussian error variable
$\varepsilon$
in \eqref{eqlinearreggroup}
and $\operatorname{df}(\lambda)$ is the number of degrees of
freedom of the selected model. Assuming that
equation~\eqref{eqlinearreggroup} holds and a differentiability
condition on
the mapping $\hat{\mathbf{y}}(\lambda)$, \citet
{2004JASAEfron}, using Stein's
theory of unbiased risk estimate [\citet{1981ASStein}], shows that
\begin{equation}
\label{eqdefdf}
\operatorname{df}(\lambda) \doteq
\frac{1}{\sigma^2}\sum_{i=1}^n \operatorname{cov}(\hat{y}_i(\lambda),
y_i) =
\mathbb{E} \biggl[ \operatorname{tr} \biggl (\frac{\partial
\hat{\mathbf{y}}(\lambda)}{\partial\mathbf{y}} \biggr) \biggr]
,
\end{equation}
where the expectation is taken with respect to $\mathbf{y}$ or,
equivalently, to
the noise~$\boldsymbol\varepsilon$.
\citet{2006JRSSYuan} proposed an approximation of the trace term
in the
right-hand side of \eqref{eqdefdf}, which is used to estimate
$\operatorname{df}(\lambda)$ for the group-Lasso:
\begin{equation}
\label{eqdfgrouplasso}
\widetilde{\operatorname{df}}_\mathrm{group}(\lambda) = \sum_{k=1}^K
\1 \bigl( \|\hatbbetagroup_{\group}(\lambda) \|>0 \bigr)
\biggl(1 +
\frac{ \|\hatbbetagroup_{\group}(\lambda) \|}
{ \|\bbetaols_{\group} \|}(p_k-1) \biggr)
,
\end{equation}
where $\1(\cdot)$ is the indicator function and $p_k$ is the number of elements
in $\group[k]$.
For orthonormal design matrices, \eqref{eqdfgrouplasso} is an unbiased
estimate of the true degrees of freedom of the group-Lasso and
\citet{2006JRSSYuan} suggest that this approximation is relevant
in more
general settings, by reporting that
``the performance of this approximate $C_p$-criterion [directly derived from~\eqref{eqdfgrouplasso}] is generally comparable with that of fivefold
cross-validation and is sometimes better.''
This approximation of $\operatorname{df}(\lambda)$ relies on the OLS
estimate and is
hence limited to setups where the latter exists and is unique. In particular,
the sample size should be larger than the number of predictors ($n \geq
p$). To
overcome this restriction,\vadjust{\goodbreak}
we suggest a more general approximation to the degrees of freedom,
based on the
ridge estimator
\begin{equation}
\label{eqbetaridge}
\hatbbetaridge(\gamma) = (\mathbf{X}^\intercal \mathbf{X} +
\gamma\mathbf{I} )^{-1} \mathbf{X}^\intercal\mathbf{y}
,
\end{equation}
which can be computed even for small sample sizes ($n<p$).
\begin{proposition} \label{propdfcooplasso} Consider the
coop-Lasso estimator $\hatbbetacoop(\lambda)$ defined by~\eqref{eqcooplassolinear}. Assuming that data are generated according to
model \eqref{eqlinearreggroup}, and that $\mathbf{X}$ is orthonormal, the
following expression of $\widetilde{\operatorname{df}}_{\mathrm
{coop}}(\lambda)$ is
an unbiased estimate of $\operatorname{df}(\lambda)$ defined in \eqref
{eqdefdf} for
the coop-Lasso fit:
\begin{eqnarray}
\label{eqdfcooplasso}
\widetilde{\operatorname{df}}_{\mathrm{coop}}(\lambda) &=& \sum_{k=1}^K
\1 \bigl( \| (\hatbbetacoop_{\group}(\lambda) )^+ \|
>0 \bigr)
\biggl(1 + \frac{p^k_+ -1}{1+\gamma}
\frac{ \| (\hatbbetacoop_{\group}(\lambda) )^+ \|}
{ \| (\hatbbetaridge_{\group}(\gamma) )^+ \|
} \biggr)
\nonumber
\\[-8pt]
\\[-8pt]
&&\hphantom{\sum_{k=1}^K}{} +
\1 \bigl( \| (\hatbbetacoop_{\group}(\lambda) )^- \|
>0 \bigr)
\biggl(1 + \frac{p^k_- -1}{1+\gamma}
\frac{ \| (\hatbbetacoop_{\group}(\lambda) )^- \|}
{ \| (\hatbbetaridge_{\group}(\gamma) )^- \|} \biggr)
,
\nonumber\hspace*{-35pt}
\end{eqnarray}
where $p^k_+$ and $p^k_-$ are respectively the number of
positive and negative entries in $\hatbbetaridge_{\group}(\gamma)$.
\end{proposition}
Proposition~\ref{propdfcooplasso} raises a practical issue regarding
the choice of a good reference $\hatbbetaridge(\gamma)$. In
our numerous simulations (most of which are not reported here), we did not
observe a high sensitivity to $\gamma$, though high values degrade
performances.
When $\mathbf{X}$ is full rank we use $\gamma=0$ (the OLS estimate) and,
correspondingly, a~vanishing $\gamma$ (the Moore--Penrose solution) when
$\mathbf{X}$ is of smaller rank.
More refined strategies are left for future works.
Section~\ref{secnumerical} illustrates that, even in nonorthonormal settings,
plugging expression \eqref{eqdfcooplasso} for the degrees of freedom
$\operatorname{df}(\lambda)$ of the coop-Lasso in BIC~\eqref{eqBICcoop} or AIC
\eqref{eqAICcoop} provides sensible model selection criteria.
As expected, BIC, which is more stringent than AIC, is better at
retrieving the
sparsity pattern of $\bbeta^\star$, while AIC is slightly better regarding
prediction error.
\section{Simulation study}\label{secnumerical}
We report here experimental results in the regression setup, with the linear
regression model \eqref{eqlinearreggroup}.
Our simulation protocol is inspired from the one proposed by
\citeauthor{1995TBreiman} (\citeyear{1995TBreiman,1996ASBreiman}) to test the
nonnegative garrote estimator,
which inspired the Lasso.
\subsection{Data generation}\label{secnumericalgenerator}
The structure of $\bbeta^\star\in\mathbb{R}^p$ is controlled
through sparsity at
coefficient and group levels.
Here we have $p=90$, forming $K=10$ groups of identical size, $p_k=9$.
All groups of parameters follow the same wave pattern: for $j\in\{
1,\ldots,9\}$,
$(\bbeta^\star_{\group})_j \propto\nu_k
( (h- |5-j | )^+ )^2$,
where $\nu_k\in\{0,1\}$ is a switch at the group level and $h\in\{
3,4,5\}$
governs the wave width, that is, the within-group sparsity, with respectively
$ |\supp_k |\in\{5,7,9\}$ nonzero coefficients in each group
included in the suppor
.
The covariates are drawn from a multivariate\vadjust{\goodbreak} normal distribution
$X \sim\mathcal{N}(\mathbf{0},\bPsi)$ with, for all $(j,j') \in
\{1,\ldots,p\}^2$, covariances $\Psi_{jj'} =
\rho^{|j-j'|}$, where $\rho\in[-1,1]$.
Finally, the response is corrupted by an error variable
$\varepsilon\sim\mathcal{N}(0,1)$ and the magnitude of the vector of
parameters
$\bbeta^\star$ is chosen to have an $R^2$ around $0.75$.
Note that the covariance of the covariates is purposely disconnected
from the
group structure. This setting may either be considered as unfair to the group
methods, or equally adverse for all Lasso-type estimators, in the sense that
none of their support recovery conditions are fulfilled when $\rho\neq0$.
Situations more or less advantageous for group methods are then
produced thanks to the parameter $h$, which determines how the support
of $\bbeta^\star$
matches the group structure.
\subsection{Results}
Model selection is performed with BIC~\eqref{eqBICcoop} for Lasso, group-Lasso
and coop-Lasso.
The estimation of the degrees of freedom for the Lasso is the number of nonzero
entries in $\hatbbetalasso(\lambda)$ [\citet{2007ASZou}].
As there is no such analytical estimate of the degrees of freedom for
the sparse
group-Lasso, we tested two alternative model selection strategies: standard
five-fold cross-validation (CV),
selecting the model with minimum cross-validation error, and the so-called
``1-SE rule'' [\citet{Breiman84}], which selects the most
constrained model whose
cross-validation error is within one standard error of the minimum.
First, we display in Figure \ref{figbreimansignalexample} an example
of the
regularization paths obtained for each method for a small training set size
($n=p/2=45$) drawn from the model with three active groups having two zero
coefficients each ($ |\supp_k |=7$, $p_k=9$) and a moderate positive
correlation level ($\rho=0.4$).
\begin{figure}
\includegraphics{520f04.eps}
\caption{Lasso, group, sparse group and coop Lasso estimates for a training
set of size $n=45$ drawn from the generation process of Section
\protect\ref{secnumericalgenerator}, with 3 active waves out of 10,
$ |\supp_k |/p_k=7/9$ and $\rho=0.4$.
Left: regularization paths, where each line type/color represents a
group of
parameters and the plain vertical line marks the model selected by the
1-SE rule for sparse group-Lasso and BIC otherwise;
right: true signal (dotted line) and estimated parameters for the selected
model (filled circles).}
\label{figbreimansignalexample}
\end{figure}
As expected, the nonzero coefficients appear one at a time along the Lasso
regularization path and groupwise for the other methods, which detect the
relevant groups early, with some coefficients kept to zero for the sparse
group-Lasso and the coop-Lasso.
The sparse group-Lasso is qualitatively intermediate between the
group-Lasso and
the coop-Lasso, setting many parameters to zero, but keeping a few negative
coefficients in the solution.
The coefficients of the model estimated by BIC or the 1-SE rule are
displayed on
the right of each path. The Lasso estimate includes some nonzero coefficients
from irrelevant groups, but is otherwise quite conservative, excluding many
nonzero parameters from its support. This conservative trend is also observed
for the group methods, which exclude all irrelevant groups.
The three group estimates mostly agree on truly important coefficients, and
differ in the treatment of the spurious negative values that are
frequent for
group-Lasso, rarer for sparse group-Lasso and do not occur for coop-Lasso.
\begin{table}
\tabcolsep=0pt
\tablewidth=330pt
\caption{Average errors, with standard deviations, on
1000 simulations from the setup described in Section~\protect\ref{secnumericalgenerator}.
Each scenario differs in the number of observations $n$ and
the number of active variables per active group
$ |\supp_k |$ ($p_k=9$).
Sparse-cv and sparse-1-se designate the sparse group-Lasso with
$\lambda$
selected by cross-validation and by the 1-SE rule,
respectively}
\label{tabbreimanconvergence}
\vspace*{-3pt}
\begin{tabular*}{330pt}{@{\extracolsep{\fill}}lcccccc@{}}
\hline
& & \textbf{Lasso} & \textbf{Group} & \textbf{Sparse-cv} &
\textbf{Sparse-1-se} & \textbf{Coop}\\
\hline
Scenario && \multicolumn{5}{c@{}}{RMSE ($\times
10^3$)} \\
$ |\supp_k |=5$ & $n=45$ & 87.1 (0.5) & 95.0 (0.5) & 82.5 (0.5) & 88.1 (0.6) & 84.2 (0.5) \\
& $n=180$ & 43.7 (0.2) & 49.1 (0.2) & 41.7 (0.2) & 44.9 (0.2) & 43.5 (0.2) \\
& $n=450$ & 28.8 (0.1) & 33.4 (0.1) & 27.2 (0.1) & 30.9 (0.1) & 29.4 (0.1) \\
$ |\supp_k |=7$ & $n=45$ & 93.0 (0.5) & 85.8 (0.5) & 79.7 (0.4) & 83.6 (0.5) & 76.8 (0.5) \\
& $n=180$ & 48.4 (0.2) & 44.5 (0.2) & 42.2 (0.2) & 43.7 (0.2) & 40.4 (0.2) \\
& $n=450$ & 31.8 (0.1) & 30.3 (0.1) & 27.7 (0.1) & 30.0 (0.1) & 27.6 (0.1) \\
$ |\supp_k |=9$ & $n=45$ & 99.2 (0.4) & 82.0 (0.5) & 81.0 (0.4) & 83.2 (0.5) & 73.7 (0.5) \\
& $n=180$ & 52.5 (0.2) & 41.9 (0.2) & 43.3 (0.2) & 43.8 (0.2) & 39.0 (0.2) \\
& $n=450$ & 34.1 (0.1) & 28.7 (0.1) & 28.8 (0.1) & 30.6 (0.1) & 27.1 (0.1) \\
[4pt]
Scenario & & \multicolumn{5}{c@{}}{Mean sign error ($\%
$)}\\
$ |\supp_k |=5$ & $n=45$ & 13.8 (0.1) & 18.3 (0.2) & 36.7 (0.4) & 16.9 (0.3) & 13.3 (0.2) \\
& $n=180$ & \hphantom{0}8.4 (0.1) & 19.3 (0.2) & 36.1 (0.4) & 10.7 (0.2) & 13.0 (0.2) \\
& $n=450$ & \hphantom{0}6.1 (0.1) & 16.7 (0.2) & 35.5 (0.4) & \hphantom{0}7.1 (0.2) & 10.3 (0.2) \\
$ |\supp_k |=7$ & $n=45$ & 18.9 (0.1) & 12.9 (0.2) & 34.6 (0.4) & 16.8 (0.3) & 10.1 (0.2) \\
& $n=180$ & 11.9 (0.1) & 12.7 (0.2) & 34.5 (0.4) & 10.5 (0.2) & \hphantom{0}9.8 (0.2) \\
& $n=450$ & \hphantom{0}8.8 (0.1) & 10.4 (0.2) & 34.9 (0.4) & \hphantom{0}7.1 (0.2) & \hphantom{0}7.7 (0.2)
\\
$ |\supp_k |=9$ & $n=45$ & 24.4 (0.1) & \hphantom{0}8.1 (0.2) & 34.2 (0.4) & 17.3 (0.3) & \hphantom{0}7.9 (0.2) \\
& $n=180$ & 15.3 (0.1) & \hphantom{0}6.3 (0.2) & 33.5 (0.4) & 10.0 (0.2) & \hphantom{0}6.7 (0.2) \\
& $n=450$ & 11.2 (0.1) & \hphantom{0}4.3 (0.1) & 32.6 (0.4) & \hphantom{0}6.0 (0.2)& \hphantom{0}4.5 (0.1)
\\
\hline
\end{tabular*}\vspace*{-3pt}
\end{table}
Table \ref{tabbreimanconvergence} provides a more objective evaluation
of the
compared methods, based on the root mean square error (RMSE) and the support
recovery (more precisely, recovery of the sign of true parameters); prediction
error (not shown) is tightly correlated with RMSE in our setup.
Regarding the relative merits of the different methods, we did not
observe a
crucial role of the number of active groups and the covariate\vadjust{\goodbreak}
correlation level
$\rho$. We report results for a true support comprising 3 groups out
of 10 and
$\rho=0.4$, with various within-group sparsity and sample size scenarios.
All estimators perform about
equally in RMSE, the sparse group-Lasso with CV having a slight
advantage over
the coop-Lasso when many zero coefficients belong to the active groups,
and the
coop-Lasso being marginally but significantly better elsewhere.
Regarding support recovery, model selection with CV leads to models
overestimating the support of parameters. The 1-SE rule, which slightly harms
RMSE, is greatly beneficial in this respect.
BIC also performs very well,
incurring a very small loss due to model selection compared to the
oracle solution picking the model with best support recovery.
The Lasso dominates all the groups methods when many zero coefficients
belong to
the active groups. Elsewhere, group methods (with appropriate model selection
criteria) perform systematically significantly better for the small
sample sizes.
The coop-Lasso ranks first or a close second among group methods in all
experimental conditions.
It thus appears as the method of choice regarding inference issues when groups
conform to the sign-coherence assumption.
\subsection{Robustness}
The robustness to violations of the sign-coherence assumption is assessed
by switching a proportion $P_\sigma$ of signs in the vector~$\bbeta
^\star$, otherwise
generated as before. The sign of the corresponding covariates are switched
accordingly, to ensure that only the coop-Lasso estimators are affected
in the
process.
Table \ref{tabbreimanrobustness} displays the coop-Lasso RMSE that degrades
gradually with the amount of perturbation, becoming eventually worse
than the
Lasso, except for full groups.
\begin{table}
\tabcolsep=0pt
\tablewidth=335pt
\caption{Average errors, with standard deviations, on
1000 simulations from the setup of Table \protect\ref{tabbreimanconvergence} with
$n=180$, perturbed by switching a proportion $P_\sigma$ of signs in
$\bbeta^\star$}
\label{tabbreimanrobustness}
\begin{tabular*}{335pt}{@{\extracolsep{\fill}}lcccccc@{}}
\hline
&\multicolumn{3}{c}{\textbf{RMSE
($\boldsymbol{\times10^3}$)}} & \multicolumn{3}{c@{}}{\textbf{Mean sign error ($\boldsymbol\%$)}} \\[-5pt]
&\multicolumn{3}{c}{\hrulefill}&\multicolumn{3}{c@{}}{\hrulefill}\\
$\boldsymbol{P_{\sigma}}$ & $ \boldsymbol{|\supp
_k |=5}$ & $ \boldsymbol{|\supp
_k |=7}$ &
{$\boldsymbol{ |\supp_k |=9}$} &
{$ \boldsymbol{|\supp_k |=5}$} &
{$\boldsymbol{ |\supp_k |=7}$} &
{$\boldsymbol{ |\supp_k |=9}$} \\
\hline
0.1 & 46.9 (0.2) & 45.3 (0.2) & 45.8 (0.2) & 15.3 (0.2) & 12.4 (0.2) &
\hphantom{0}8.8 (0.2) \\
0.2 & 49.5 (0.3) & 48.9 (0.2) & 48.7 (0.2) & 17.8 (0.2) & 14.3 (0.2) &
\hphantom{0}9.8 (0.2) \\
0.3 & 51.0 (0.3) & 50.4 (0.3) & 50.4 (0.2) & 19.3 (0.2) & 14.8 (0.2) &
10.3 (0.2) \\
0.4 & 51.6 (0.2) & 51.0 (0.2) & 50.2 (0.2) & 19.7 (0.2) & 14.8 (0.2) &
\hphantom{0}9.8 (0.2) \\
0.5 & 52.3 (0.3) & 51.3 (0.2) & 50.8 (0.2) & 20.0 (0.2) & 14.6 (0.2) &
\hphantom{0}9.3 (0.2) \\
\hline
\end{tabular*}
\end{table}
Regarding sign error, for small proportions of sign flip, the
coop-Lasso stays at
par with either Lasso or group-Lasso (see Table \ref{tabbreimanconvergence}),
but it eventually becomes significantly worse than both of them in most
situations. Thus, if the sign-coherence assumption is not firmly grounded,
either group-Lasso or its sparse version seem to be better options:
coop-Lasso only remains a second-best choice when there are less than
10\%
of sign mismatches within groups.
\section{Illustrations on real data}\label{secdata}
This section illustrates the applicability of the coop-Lasso on two
types of
predictors, that is, categorical and continuous covariates.
The first proposal may be widely applied to ordered categorical variables;
the second one is specific to microarray data, but should apply more generally
when groups of variables are produced by clustering.
In the first application, each group is formed by a set of variables
coding an ordered categorical variable.
Ordinal data are often processed
either by omitting the order property, treating them as nominal, or by
replacing each level with a
prescribed value, treating them as quantitative.
The latter procedure, combined with generalized linear regression,
leads to
monotone mapping from levels to responses.
Section \ref{secdataordinal} describes how coop-Lasso can bias the estimate
toward monotone mappings using a categorical treatment of ordinal variables.
In the second application of Section \ref{secdatacontinuous}, the
groups are
formed by continuous variables that are redundant noisy measurements (probe
signals) pertaining to a\vadjust{\goodbreak} common higher-level unobserved variable (gene
activity).
Sign-coherence is expected here, since each measurement should be positively
correlated with the activity of the common unobserved variable.
A similar behavior should also be anticipated when groups of
variables are formed by a clustering preprocessing step based on the Euclidean
distance, such as $k$-means or average linkage hierarchical clustering
[\citet{1998PNASEisen}; \citet{2006BSPark}; \citet
{2007BMCMa}].
\subsection{Monotonicity of responses to ordinal covariates}\label
{secdataordinal}
Monotonicity is easily dealt with by transforming ordinal covariates into
quantitative variables, but this approach is arbitrary and subject to many
criticisms when there is no well-defined numerical difference between levels,
which often lacks even for interval data when the lower or the upper
interval is
not bounded [\citet{Gertheiss09}].
Hence, the categorical treatment is often preferred, even if it fails
to fully
grasp the order relation.
The Lasso, group-Lasso or fused-Lasso have been applied to
the categorical treatment of ordinal features, with the aim to select
variables or aggregate adjacent levels [see \citet{Gertheiss10}
and references
within].
The coop-Lasso is used here to make a stronger usage of the order
relationship, by biasing
the mapping from levels to the response variable toward monotonic solutions.
Note that our proposal does not impose monotonicity and neither does it
prescribe an order (although several variations would be possible here).
In these respects, we depart from the approaches
imposing hard constraints on regression coefficients [\citet
{Rufibach20101442}].
\subsubsection{Methodology}
When not treated as numerical, ordinal variables are often coded by a
set of
variables that code differences between levels. Several types of
codings have
been developed in the ANOVA setting, with relatively little impact in the
regression setting, where the so-called dummy codings are intensively used.
Indeed, least squares fits
are not sensible to coding choices provided there is a one-to-one
mapping from
one to the other, so that codings only matter regarding the direct
interpretation
of regression coefficients.
However, codings evidently affect the solution in penalized
regression, and we will use here specific codings to penalize targeted
variations.
In order to build a monotonicity-based penalty, we simply use contrasts
that compare
two adjacent levels. An example of these contrasts is displayed in
Table~\ref{tabapplicationcontrasts}, with the corresponding codings,
known as
backward difference codings, which are simply obtained by solving a linear
system [\citet{Serlin85}].
\begin{table}
\tabcolsep=0pt
\tablewidth=205pt
\caption{Contrasts and codings for comparing
the adjacent levels of a covariate with 4 levels}\label{tabapplicationcontrasts}
\begin{tabular*}{205pt}{@{\extracolsep{\fill}}ld{2.0}d{2.0}d{2.0}d{4.0}d{4.0}d{4.0}@{}}
\hline
\textbf{Level} & \multicolumn{3}{c}{\textbf{Contrasts}} &
\multicolumn{3}{@{}c}{\textbf{Codings}} \\ \hline
0 & -1 & 0 & 0 & -3/4 & -1/2 & -1/4 \\
1 & 1 & -1 & 0 & 1/4 & -1/2 & -1/4 \\
2 & 0 & 1 & -1 & 1/4 & 1/2 & -1/4 \\
3 & 0 & 0 & 1 & 1/4 & 1/2 & 3/4 \\
\hline
\end{tabular*}
\end{table}
Note that several codings are possible for the contrasts given in
Table~\ref{tabapplicationcontrasts}. They differ in the definition of a
global reference level, whose effect is relegated to the intercept. As
we do not
penalize the intercept here, the particular choice has no outcome on the
solution.
Irrespective of the coding, group penalties act as a selection tool for factors,
that is, at variable level [\citet{2006JRSSYuan}].
On top of this, the sparse group penalty\vadjust{\goodbreak} usually presents the ability to
discard a level. With difference codings, some increments between adjacent
levels may be set to zero, that is, levels may be fused [\citet
{Gertheiss10}].
With the coop-Lasso penalty, all increments are urged to be sign-coherent,
thereby favoring monotonicity. As a side effect, level fusion may also be
obtained.
\subsubsection{Experimental setup}
We illustrate the approach on the Statlog ``German Credit'' data set [available
at the UCI machine learning repository, \citet{Frank10}], which gathers
information about people classified as low or high credit risks.
This binary response requires an appropriate model, such as logistic regression.
The coop-Lasso fitting algorithm is easily adaptable to generalized linear
models, following exactly the structure provided in Algorithm
\ref{algoactiveconstraint}, where the appropriate likelihood function replaces
the sum of square residuals in Step \ref{itemalgostep1}.
All quantitative variables
are used for the analysis, but we focus here on the regression
coefficients of
four variables, encoded as integers or nominal in the Statlog project, which
seem better interpreted as ordered nominal, namely:
\texttt{history}, with 4 levels describing the ability to pay back
credits in the past and now;
\texttt{savings}, with 4 levels giving the balance of the saving
account in currency intervals;
\texttt{employment}, with 5 levels reporting the duration of the
present employment in year intervals; and
\texttt{job}, with 4 levels representing an employment qualification scale.
Two other variables, related to the checking account status and
property, were
also encoded as nominal, but are not described here in full details
since they do not show distinct qualitative behaviors between methods.
We excluded from the ordinal variables categories merging two subcategories
possibly corresponding to different ranks,
such as
``critical account/other credits existing (not at this bank)'' in
\texttt{history}, or ``unknown/no savings account'' in \texttt{savings}.
For simplicity, we suppressed the corresponding examples, thus ending
with a~total of 330 observations, split into three equal-size learning,
validation and test sets. We estimate the logistic regression
coefficients on
the learning set, perform model selection from deviance or
misclassification error on
the validation set, and finally keep the test set to estimate
prediction performances.\looseness=1
\subsubsection{Results}
The performances of the three group methods are identical, either
evaluated in terms of\vadjust{\goodbreak}
deviance, classification error rate or weighted misclassification (unbalanced
misclassification losses are provided with the data set).
The regression coefficients differ, however, as shown in Figure
\ref{figgermancreditdatapath} displaying the regularization paths for all
methods.
\begin{figure}
\includegraphics{520f05.eps}
\caption{Regularization paths for four ordinal covariates (history, savings,
job and employment) for the group, coop, and sparse group-Lasso on
the contrast coefficients obtained from backward difference coding
(top left, top right and bottom left, respectively). The
transcription of contrasts to levels is also displayed for coop-Lasso
(bottom right). The vertical lines mark the model selected by
cross-validation on the validation set, for different criteria:
deviance (plain), misclassification rate (dashed), and weighted
misclassification error (dotted).}
\label{figgermancreditdatapath}
\end{figure}
Recall that we only represent the ordinal covariates \texttt{history},
\texttt{savings}, \texttt{employement} and \texttt{job}. Each
coefficient represents
the increment between two adjacent levels, with positive and negative values
resulting in an increase and decrease, respectively.
Monotonicity with respect to all levels is reached if all the values
corresponding to a factor are nonnegative or nonpositive.
We also provide an alternative view of the coop-Lasso path, with the overall
effects corresponding to levels, obtained by summing up the increments.
Most factors are not obviously amenable to quantitative coding since
there is no natural distance between levels, but we, however, underline that
using the usual quantitative transformation with equidistant values
followed by
linear regression would correspond here to identical increments between levels.
Obviously, all displayed solutions radically contradict this linear
trend hypothesis.
Our three solutions differ regarding monotonicity, which is almost never
observed along the group-Lasso regularization path. The sparse
group-Lasso paths
have long sign-coherent sections, where group-Lasso infers slight
wiggles. These
sections extend further with the coop-Lasso.
However, as the coop penalty goes to zero, sign-coherence is no longer
preserved, and all methods eventually reach the same solution.
The sparse group and the coop-Lasso set some increments to zero,
leading to the
fusion of adjacent levels that should be welcomed regarding interpretation.
The solutions tend to agree on these fusions on long sections of the
paths, with
some additional fusions of the sparse group-Lasso when slight monotonic
solutions are provided by the coop-Lasso (see \texttt{employment},
levels 2 and
3, and \texttt{savings} levels 1 and 2).
These fusions are perceived more directly on the coop-Lasso path of effects,
displayed in the bottom right of Figure~\ref{figgermancreditdatapath}, where
the effect of each level is displayed directly.
\subsection{Robust microarray gene selection}\label{secdatacontinuous}
Most studies on response to che\-motherapy have considered breast cancer
as a single homogeneous entity. However, it is a complex disease
whose strong heterogeneity should not be overlooked. The data set
proposed by \citet{2006JCOhess} consists in gene expression
profiling of patients treated with chemotherapy prior to surgery,
classified as presenting either a pathologic complete response (pCR) or a~residual disease (not-pCR). It records the signal of 22,269
probes\footnote
Actually, the data set reports the average signal in probe sets, which
are a
collection of probes designed to interrogate a given sequence. In this paper
the term ``probe'' designates Affymetrix probe sets to avoid confusion
with the
group structure that will be considered at a higher level.}
examining the human genome, each probe being related to a unique gene.
Following \citet{2011JSFDSjeanmougin}, we restrict our analysis
to the basal
tumors:
for this particular subtype of breast cancer, clinical and pathologic features
are homogeneous in the data set, whereas the response to chemotherapy is
balanced, with 15 tumors being labeled pCR and 14 not-pCR.
This setup is thus propitious to the statistical analysis of response to
chemotherapy from the sole activity of genes.
\subsubsection{Methodology}
The usual processing of microarray data relies on probe
measurements\vadjust{\goodbreak}
that are related to genes in the final interpretation of
the statistical analysis. Here we would like to take a different
stance, by
gathering all the measurements associated to gene entities at an early
stage of
the statistical inference process.
As a matter of fact, we typically observe that some probes related to
the very
same gene have different behaviors.
Requiring a consensus at the gene level supports biological coherence, thus
exercising caution in an inference process where statistically
plausible explanations are numerous, due to the noisy probe signals and
to the
cumbersome $n \ll p$ setup (here $n=29$ and $p=22\mbox{,}269$).
Since the probes related to a given gene relate to sequences that are
predominantly cooperating, the sign-coherence assumed by the coop-Lasso is
particularly appropriate to improve robustness to the measurement noise
and to
encourage biologically plausible solutions.
Our protocol includes a preselection of probes that facilitates the
analysis for
the nonadaptive penalization methods compared here, and also provides an
assessment of the benefits of adding seemingly less relevant
probes into the statistical analysis. We proceed as follows:
\begin{itemize}
\item select a restricted number $d$ of probes from classical differential
analysis, where probes are sorted by increasing $p$-values;
\item determine the genes associated to these $d$ probes, retrieve all the
probes related to these genes, and select the corresponding $p$ probes,
$p \ge d$,
regardless of their signal;
\item fit a model with group penalties where groups are defined by genes.
\end{itemize}
\subsubsection{Experimental setup}
We select the first $d=200$ most differentiated probes, as identified
by the analysis of \citet{2011JSFDSjeanmougin}, on the 22,269 probes
for the $n=29$ patients with basal tumor. These 200 probes correspond
to 172 genes, themselves associated to $p=381$ probes on the
microarray as a whole, with 1 to 13 probes per gene. We clearly enter
the high-dimensional setup with $p > 13 \times n$.
All signals are
normalized to have a unitary within-class variance. We compare
then the Lasso on the $d=200$ most differentiated probes, with the
Lasso and group, sparse group and coop Lasso on the $p=381$ probes.
All fits are produced with our code (available at
\texttt{\href{http://stat.genopole.cnrs.fr/logiciels/scoop}{http://stat.genopole.cnrs.fr/}
\href{http://stat.genopole.cnrs.fr/logiciels/scoop}{logiciels/scoop}}).
Well-motivated analytical model selection criteria are not available
today for Lasso-type penalties beyond the regression setup. Here,
model selection is carried out by 5-fold cross-validation: we
evaluate the $\mathrm{CV}$ error for each method with the same block
partition using either the binomial deviance or the unweighted
classification error.
\subsubsection{Results}
The 5-folds $\mathrm{CV}$ scores, either based on deviance or
misclassification losses,\vadjust{\goodbreak} are reported for each estimation method in
Table \ref{tabcvresults}, which also displays the number of selected groups
and features for the models selected by minimizing the CV score.
\begin{table}
\tabcolsep=0pt
\caption{$\mathrm{CV}$ scores for misclassification error and
binomial deviance on the basal tumor data.
The minimizer of $\mathrm{CV}$ for misclassification and deviance
are respectively denoted by $\lambdaclass$ and~$\lambda^{\mathrm{dev}}$; the number of selected groups and features
respectively refers to genes and probes}
\label{tabcvresults}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccccc@{}}
\hline
& & \textbf{Probes} & \textbf{Lasso} & \textbf{Group} &
\textbf{Sparse} & \textbf{Coop} \\
\hline
Model selection rule &&
\multicolumn{5}{c@{}}{CV score $\times100$
(standard error)} \\
Classification
& $\lambdaclass$ & 10.3 (5.8)\hphantom{0} & 6.9 (4.9) & \hphantom{0}3.4 (3.5) & 3.4 (3.5)
& \hphantom{0}3.4 (3.5) \\
Deviance &
$\lambda^{\mathrm{dev}}$ & 76.5 (37.6) & 67.2 (32.3) & 13.7 (8.1) & 20.5 (10.0) & 13.8 (7.9) \\[4pt]
Model selection rule &&
\multicolumn{5}{c@{}}{\# selected groups (features)}\\
Classification
& $\lambdaclass$ & 17 (17) & 16 (17) & 11 (15) & 14 (21)
& \hphantom{0}9 (11) \\
Deviance
& $\lambda^{\mathrm{dev}}$ & 19 (19) & 17 (18) & 13 (21) & 16 (26)
& 14 (18) \\
\hline
\end{tabular*}
\end{table}
Expanding the set of probes from $d$ to $p$ slightly improves the performances
of the Lasso, and considerable further progresses are brought by all group
methods, which misclassify about 1 patient among the 29 and quarter deviance
scores.\footnote
A note of caution regarding performances: scores comparisons are
fair here, in the sense that the $\mathrm{CV}$ scores are optimized with
respect to a single parameter $\lambda$, whose role is analog for
all. Additional simulations (not reported here) show that, for all group
methods, the
${\mathrm{CV}}$ error is stable with respect to the random choice of
folds and that the $\mathrm{CV}$ curves are smooth around their
minima. However, the minimizers of
$\mathrm{CV}$ are biased estimates of out-of-sample scores, and the
representativeness of their observed difference
can be questioned.}
As expected, less genes are selected by group methods;
the difference is more important for the minimizers of the
misclassification score, and, among those, for the group-Lasso and coop-Lasso
that comply more stringently to the group structure.
These observations indicate that the group structure defined by genes provides
truly useful guidelines for inference.
The sparsity numbers differ among the group methods, coop-Lasso
selecting as
many genes as group-Lasso and fewer probes, and sparse group-Lasso retaining
slightly more genes and probes.
A more detailed picture is provided in Figure~\ref{figprobes}, which shows
the regression coefficients for the three group estimators adjusted on
the whole
data set with their respective $\lambdaclass$ values.
\begin{figure}
\includegraphics{520f06.eps}
\caption{Logistic regression coefficients attached to each probe for
group, sparse group and coop Lasso. Each marker (color and
symbol) designates the gene associated to the probe:
\textsc{rnps1}
\mbox{\emph{(}\protect\includegraphics{520i01.eps}\emph{)}},
\textsc{msh6}
\mbox{\emph{(}\protect\includegraphics{520i02.eps}\emph{)}},
\textsc{prps2}
\mbox{\emph{(}\protect\includegraphics{520i03.eps}\emph{)}},
\textsc{h1fx}
\mbox{\emph{(}\protect\includegraphics{520i04.eps}\emph{)}},
\textsc{mfge8}
\mbox{\emph{(}\protect\includegraphics{520i05.eps}\emph{)}},
\textsc{sulf1}
\mbox{\emph{(}\protect\includegraphics{520i06.eps}\emph{)}},
\textsc{rnf115}
\mbox{\emph{(}\protect\includegraphics{520i07.eps}\emph{)}},
\textsc{rnf38}
\mbox{\emph{(}\protect\includegraphics{520i08.eps}\emph{)}},
\textsc{thnsl2}
\mbox{\emph{(}\protect\includegraphics{520i09.eps}\emph{)}}
and \textsc{edem3}
\mbox{\emph{(}\protect\includegraphics{520i10.eps}\emph{)}}.}
\label{figprobes}
\end{figure}
Among the three methods, a~total of 15 groups (i.e., genes) are selected.
For readability, we only represent the 10 leading groups of regression
coefficients (according to their average norm).
We first oberve that the magnitude of coefficients differs for each
method, the
coop-Lasso having the smallest one. In fact, there is a wide range of
$\lambdaclass$ values for which the miclassification score is minimal\vadjust{\goodbreak}
for the
coop-Lasso, enabling to choose a highly penalized solution without affecting
accuracy.
The magnitude apart, the group methods have qualitatively the same
behaviors for
all unitary groups but one, with \textsc{thnsl2}
(\includegraphics{520i09.eps}) being set to zero
by the
coop-Lasso.
The same patterns are observed for two other groups,
\textsc{rnps1} ({\includegraphics{520i01.eps}}) and
\textsc{edem3} ({\includegraphics{520i10.eps}}), whose
regression coefficients are consistently estimated to be sign-coherent.
Then, \textsc{sulf1} ({\includegraphics{520i06.eps}}),
though being estimated sign-coherent by the sparse group-Lasso, is
excluded from
the support of the group and coop Lasso.
Finally, \textsc{msh6} ({\includegraphics{520i02.eps}}),
estimated as sign incoherent with the two groups methods, is excluded
from the
support for the coop-Lasso.
Overall, the probe enrichment scheme we propose here leads to considerable
improvements in prediction performance. This better statistical
explanation is
obtained without impairing interpretability, since sign-coherence is actually
often satisfied by all methods and strictly enforced by the coop-Lasso.
As often in this type of study, several methods provided similar
prediction performances, but the explanation provided by the coop-Lasso is
simpler, both from a statistical and from a biological viewpoint. Note
that the
coefficient paths (not shown) diverge early between the group and coop methods,
so that the above-mentioned discrepancies are not simply due to model selection
issues.
As a final remark, we observed qualitatively similar behaviors when
the initial number of probes $d$ ranged from 10 to 2000.
For $d \leq1000$, the group methods always performed best, with
approximately identical
classification errors, the group-Lasso and coop-Lasso slightly
dominating the
sparse group-Lasso in terms of deviance. With larger initial sets of probes,
the enrichment procedure becomes less efficient, and all methods provide
similar decaying results. The chosen setup displayed here, with
$d=200$, leads to the
smallest classification error for all methods, and was chosen for being
representative of the most interesting regime.\vadjust{\goodbreak}
\section{Discussion}\label{secdiscussion}
The \emph{coop-Lasso} is a variant of the \emph{group-Lasso} that was
originally proposed in the context of multi-task learning, for
inferring related networks with Gaussian Graphical Models
[\citet{2010SCChiquet}]. Here we develop its analysis in the linear
regression setup and demonstrate its value for prediction and inference with
generalized linear models.
Along with this paper we provide an implementation of the fitting
algorithm in the
\texttt{R} package \texttt{scoop}, which makes this new penalized estimate
publicly available for linear and logistic regression (the \emph
{coop-Lasso} for
multiple network inference is also available in the \texttt{R} package
\texttt{simone}).
The coop-Lasso differs from the group-Lasso and sparse group-Lasso
[Fried\-man, Hastie and
Tibshirani (\citeyear{2010preprintFriedman})] by the assumption that the group
structure is \emph{sign-coherent}, namely, that groups gather either
nonpositive, nonnegative or null parameters, enabling the recovery of various
within-group sign patterns (positive, negative, null, nonpositive,
nonnegative, nonnull).
This flexibility greatly reduces the incentive to drive within-group
sparsity with an additional parameter that later
leads to an unwieldy model selection step.
However, the relevance of the sign-coherence assumption should be firmly
established since it plays an essential role in the performance of
coop-Lasso compared to the sparse group-Lasso.
Under suitable irrepresentable conditions, the proposed penalty leads
to consistent model selection, even when the true sparsity pattern
does not match the group structure. When the groups are sign-coherent
the coop-Lasso compares favorably to the group-Lasso, recovering the
true support under the mildest assumptions.
We present an approximation of the effective degrees of freedom of the
coop-Lasso which, once plugged into AIC or BIC, provides a fast way to select
the \mbox{tuning} parameter in the linear regression setup.
We provide empirical results demonstrating the capabilities of the
coop-Lasso in
terms of prediction and parameter selection, with BIC performing very well
regarding support recovery even for small sample sizes.
We illustrate the merits of the coop-Lasso applied to the analysis to ordinal
and continuous predictors. With
an apposite coding, such as forward or backward
difference coding, the sign-coherence assumption is transcribed
in a monotonicity assumption, which does not require to stipulate the
usual and controversial mapping from levels to quantitative variables.
Finally, the application to genomic data opens a vast potential field
of great
practical interest for this type of penalty, both in terms of
prediction and
interpretability.
Our forthcoming investigations will aim at substantiating this ambition by
conducting large scale experiments in this application domain.
\begin{appendix}
\section*{Appendix: Proofs}\label{appm}
\subsection{\texorpdfstring{Proof of Lemma~\protect\ref{lemsubdifferential}}{Proof of Lemma 1}}
Let us use
${\cal T}_k$ as a shorthand for ${\cal S}_k(\bbeta)$, \citet
{2010SCChiquet} show that the
subdifferential $\btheta$ obey the following conditions:
\begin{subequation}
\label{eqsubgradient}
\begin{eqnarray}\label{eqsubgradienta}
&\displaystyle \max (\norm{\btheta_{\group}^+},\norm{\btheta_{\group
}^-} )\leq w_k
\qquad \mbox{if } \bbeta_{\group} = \mathbf{0} ,&
\\\label{eqsubgradientb}
&\displaystyle \btheta_{{\cal T}_k} = \frac{w_k \bbeta_{{\cal T}_k}}{\norm{\bbeta_{{\cal
T}_k}}},\qquad
\norm{\btheta_{{\cal T}_k^c}^-} \leq w_k, \qquad
\norm{\btheta_{{\cal T}_k^c}^+} = 0&\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{\mbox{if } \|\bbeta_{\group}^+\|>0, \|\bbeta
_{\group}^-\|=0 ,}
\\\label{eqsubgradientc}
&\displaystyle \btheta_{{\cal T}_k} = \frac{w_k \bbeta_{{\cal T}_k}}{\norm{\bbeta_{{\cal
T}_k}}},\qquad
\norm{\btheta_{{\cal T}_k^c}^+} \leq w_k, \qquad
\norm{\btheta_{{\cal T}_k^c}^-} = 0
&\nonumber
\\[-8pt]
\\[-8pt] \eqntext{\mbox{if } \|\bbeta_{\group}^-\|>0, \|\bbeta
_{\group}^+\|=0 ,}
\\
\label{eqsubgradientd}
&\displaystyle \forall j\in\group \qquad \btheta_{j} = w_k \bbeta_{j} \norm{\sign(\beta_j)\bbeta}^{-1}
&\nonumber
\\[-8pt]
\\[-8pt] \eqntext{\mbox{if } \|\bbeta_{\group}^-\|>0, \|\bbeta
_{\group}^+\|>0 .}
\end{eqnarray}
\end{subequation}
We thus simply have to prove the equivalence of conditions
\eqref{eqsubgradientcompact} and \eqref{eqsubgradient} for all
$\bbeta_{\group}$
values.
For $\bbeta_{\group}=\mathbf{0}$, \eqref{eqsubgradientcompact} reads
\begin{equation}
\norm{\btheta_{\group}^+} \leq w_k \quad \mbox{and} \quad
\norm{\btheta_{\group}^-} \leq w_k
,
\label{eqsubgradientabis}
\end{equation}
which is equivalent to \eqref{eqsubgradienta}.
For $\bbeta_{\group}\neq\mathbf{0}$, the equalities for $\btheta
_{{\cal T}_k}$ in
\eqref{eqsubgradientb}--\eqref{eqsubgradientd} are equivalent to
\eqref{eqsubgradientcompactb}, thus setting the equivalence between
\eqref{eqsubgradientcompact} and \eqref{eqsubgradient} for all nonzero
coefficients.
For $\bbeta_{{\cal T}_k^c}$, let us consider the case~\eqref{eqsubgradientb},
where all nonzero parameters within group $k$ are positive.
The first equation of~\eqref{eqsubgradientb} implies that $\norm{\btheta_{{\cal T}_k}^+} = w_k$
and $\norm{\btheta_{{\cal T}_k}^-} = 0$.
Hence, $\norm{\btheta_{{\cal T}_k^c}^-} \leq w_k$ and
$\norm{\btheta_{{\cal T}_k^c}^+}=0$ imply \eqref
{eqsubgradientabis}, so that
\eqref{eqsubgradientb} implies~\eqref{eqsubgradientcompact}.
The contraposition is also easy to check. From \eqref{eqsubgradientcompactb},
when all coefficients are positive, we have that $\norm{\btheta
_{{\cal T}_k}^-}=0$
and $\norm{\btheta_{{\cal T}_k}^+} = w_k$.
Then, this implies that \eqref{eqsubgradientcompacta} reads
\[
\norm{\btheta_{{\cal T}_k^c}^-} \leq w_k
\quad \mbox{and} \quad
\norm{\btheta_{{\cal T}_k^c}^+}=0
,
\]
which defines $\btheta_{{\cal T}_k^c}$ in \eqref{eqsubgradientb}. The
proof is similar
for \eqref{eqsubgradientc} where all nonzero parameters within group
$k$ are
positive.
\subsection{\texorpdfstring{Proof of Proposition~\protect\ref{propdfcooplasso}}{Proof of Proposition 1}}
We assume here that $\mathbf{X}^\intercal \mathbf{X} = \mathbf{I}_p$.
We introduce the ridge estimator in the computation of the trace in
equation~\eqref{eqdefdf}, through the chain rule, yielding an unbiased
estimate of $\operatorname{df}$:
\begin{eqnarray*}
\widetilde{\operatorname{df}}_{\mathrm{coop}}(\lambda)
&= &\operatorname{tr} \biggl(\frac{\partial\hat{\mathbf{y}}(\lambda
)}{\partial\mathbf{y}} \biggr)
= \operatorname{tr} \biggl(
\frac{\partial\mathbf{X}^{\intercal}\hatbbetacoop(\lambda)}
{\partial\hatbbetaridge(\gamma)}
\,\frac{\partial\hatbbetaridge(\gamma)}{\partial\mathbf{y}}
\biggr) \\
&= & \frac{1}{1 + \gamma} \sum_{k=1}^K \sum_{j\in\group} \frac
{\partial
\hatbetacoop_j(\lambda)}{\partial\hatbetaridge_j(\gamma)}
,
\end{eqnarray*}
where the last equation derives from the definition \eqref
{eqbetaridge} of the
ridge estimator with regularization parameter $\gamma$. Then, the
expression of
the coop-Lasso as a function of the ridge regression estimate is simply obtained
from equation~\eqref{eqcooportho}, using that, in the orthonormal
case, we have $\hatbbetaols= (1+\gamma)
\hatbbetaridge(\gamma)$. Dropping the reference to $\lambda$ and
$\gamma$ that
is obvious from the context, we have, $\forall k \in\{1,\ldots,K\}$ and
$\forall j\in\group$,
\begin{equation}
\label{eqcoopridgeortho}
\hatbeta_j^{\mathrm{coop}} = \biggl (
1-\frac{\lambda
w_k}{(1+\gamma) \|\bvarphi_j(\hatbbetaridge_{\group}) \|}
\biggr)^{ +}
(1+\gamma) \hatbeta_j^{\mathrm{ridge}}
.
\end{equation}
Then, for $j\in\group$, routine differentiation gives
\begin{eqnarray*}
\frac{1}{1+\gamma}\,\frac{\partial\hatbetacoop_j}{\partial
\hatbetaridge_j} &= &
\1 (\|\hatbetacoop_j\|>0 ) \\
&&{} \times \biggl(1 -
\frac{\lambda w_k} {(1+\gamma)} \biggl(
\frac{1}{\|\bvarphi_j(\hatbbetaridge_{\group})\|} -
\frac{(\hatbetaridge_j)^2}{\|\bvarphi_j(\hatbbetaridge_{\group})\|^3}
\biggr)
\biggr)
.
\end{eqnarray*}
The summation over the positive and negative elements of $\group$
reduces to two
terms
\begin{eqnarray*}
&&\frac{1}{1+\gamma} \sum_{j\in\group}\frac{\partial\hatbetacoop
_j}{\partial
\hatbetaridge_j}\\
&& \qquad = \1 \bigl(\|(\hatbbetacoop_{\group})^+\|>0 \bigr)
\biggl(p^k_+ -
\frac{\lambda w_k}{1+\gamma}
\frac{(p^k_+ -1)}{\|(\hatbbetaridge_{\group})^+\|} \biggr)
\\ && \qquad \quad {}+
\1 \bigl(\|(\hatbbetacoop_{\group})^-\|>0 \bigr) \biggl(p^k_- -
\frac{\lambda w_k}{1+\gamma}
\frac{(p^k_- -1)}{\|(\hatbbetaridge_{\group})^-\|} \biggr) \\
&& \qquad = \1 \bigl(\|(\hatbbetacoop_{\group})^+\|>0 \bigr) +
\biggl (1 - \frac{\lambda w_k}{(1+\gamma)\|(\hatbbetaridge_{\group
})^+\|} \biggr)^+ (p^k_+ -1)
\\ && \qquad \quad {}+ \1 \bigl(\|(\hatbbetacoop_{\group})^-\| >0 \bigr) + \biggl (1 -
\frac{\lambda w_k}{(1+\gamma)\|(\hatbbetaridge_{\group})^-\|}
\biggr)^+(p^k_- -1)
.
\end{eqnarray*}
From \eqref{eqcoopridgeortho}, we have, $\forall k \in\{1,\ldots,K\}
$ and $\forall j\in\group ,
$
\[
\biggl(
1-\frac{\lambda w_k}
{(1+\gamma) \|\bvarphi_j(\hatbbetaridge_{\group}) \|}
\biggr)^{ +}
= \frac{1}{1+\gamma} \frac{ \|\bvarphi_j(\hatbbetacoop
_{\group}) \|}{ \|\bvarphi_j(\hatbbetaridge_{\group
}) \|}
,
\]
which is used twice to simplify the previous expression. Summing over all
groups concludes the proof.
\subsection{\texorpdfstring{Proof of Theorem \protect\ref{thmsupportconsistency}}{Proof of Theorem 2}}
Our asymptotic results are established on the scaled problem~\eqref
{eqcooplassonormalized}.\vadjust{\goodbreak}
We then follow the three steps proof technique proposed by \citet
{2007JRSSYuan} for
the Lasso and also applied by \citet{2008JMLRBach} for the group-Lasso:
\begin{longlist}[(3)]
\item[(1)] restrict the estimation problem to the true support;
\item[(2)] complete this estimate by 0 outside the true support;
\item[(3)] prove that this artificial estimate satisfies optimality
conditions for the original coop-Lasso problem with probability
tending to 1.
\end{longlist}
Then, under (A2), the solution is unique, leading to the conclusion
that the
coop-Lasso estimator is equal to this artificial estimate with
probability tending to 1, which ends the proof. Note, however, a slight
yet important difference along the discussion: since we authorize
divergences between the group structure $\{\group\}_{k=1}^K$ and the true
support $\supp$, the irrepresentable conditions (A4)--(A5) for the
coop-Lasso cannot
be expressed simply in terms of coop-norms [as it is done with the
group-norm in \citet{2008JMLRBach}].
We will see that this does not impede the development of the proof.
As a first step, we prove two simple lemmas. Lemma
\ref{lemsupportconsistency} states that the coop-Lasso estimate,
restricted on
the true support $\supp$, is consistent when $\lambda_n \to0$. Lemma
\ref{lemUseIrrepresentableCond} provides the basis for the inequalities
\eqref{thICnoPnoN} and \eqref{thICPorN} that express our irrepresentable
conditions.
\begin{lemma} \label{lemsupportconsistency} Assuming \textup{(A1)--(A3)}, let
$\tildebbeta_\supp^n$ be the unique minimizer of the regression
problem restricted to the true support $\supp$:
\[
\tildebbeta_\supp^n = \argmin_{\mathbf{v} \in\Rset^{|\supp|}}
\frac{1}{2} \|\mathbf{y} - \mathbf{X}_{\centerdot\supp}\mathbf
{v}\|_n^2 +
\lambda_n \sum_{k\dvtx\supp_k \neq\varnothing} w_k (\|\mathbf{v}_{\supp
_k}^+\| +
\|\mathbf{v}_{\supp_k}^-\|)
,
\]
where $\|\cdot\|_n = \|\cdot\|/n$ denotes the empirical norm.
If $\lambda_n \rightarrow0$, then $\tildebbeta_\supp^n \inprob
\bbeta^\star_{\supp}.$
\end{lemma}
\begin{pf}
This lemma stems from standard results of M-estimation
[\citet{1998CUPVaart}].
Let $\boldsymbol\varepsilon= \mathbf{y} - \mathbf{X}\bbeta^\star
$, and write
$\bPsi^n = \mathbf{X}^\intercal\mathbf{X}/n$.
If $\lambda_n \rightarrow0$, then under (A1)--(A2), for any $\mathbf{v}
\in
\Rset^{|\supp|}$
\begin{eqnarray*}
Z_n(\mathbf{v}) & =&
\frac{1}{2} \|\mathbf{y} -
\mathbf{X}_{\centerdot\supp}\mathbf{v}\|_n^2 + \lambda_n \sum
_{k\dvtx \supp_k \neq
\varnothing} w_k (\|\mathbf{v}_{\supp_k}^+\| + \|\mathbf{v}_{\supp
_k}^-\|) \\
&=& \frac{1}{2}
(\bbeta^\star_{\supp}-\mathbf{v})^\intercal\bPsi_{\supp\supp}^n
(\bbeta^\star_{\supp}-\mathbf{v}) -
\frac{1}{n} \boldsymbol\varepsilon^\intercal\mathbf{X}_{\centerdot
\supp}(\bbeta^\star_{\supp}-\mathbf{v}) +
\frac{\boldsymbol\varepsilon^\intercal \boldsymbol\varepsilon
}{2n} \\
&&{} + \lambda_n \sum_{k,
\supp_k \neq\varnothing} w_k (\|\mathbf{v}_{\supp_k}^+\| + \|\mathbf
{v}_{\supp_k}^-\|)
\end{eqnarray*}
tends in probability to
\[
Z(\mathbf{v}) =
\tfrac{1}{2}(\bbeta^\star_{\supp}-\mathbf{v})^\intercal\bPsi
_{\supp\supp} (\bbeta^\star_{\supp}-\mathbf{v}) +
\tfrac{1}{2}\sigma^2.
\]
It follows from the strict convexity of $Z_n$ that $\argmin Z_n(\mathbf{v})
\inprob\argmin Z(\mathbf{v}) = \bbeta^\star_{\supp}$ [\citet
{2000ASKnight}], which ends the
proof.\vadjust{\goodbreak}
\end{pf}
\begin{lemma} \label{lemUseIrrepresentableCond}
Consider a sequence of random variables $S_n$ such that
$S_n \inprob S$. Suppose there exists $\delta>0$ such that for
a given norm
$\mu$ the limit $S$ is bounded away from 1:
\[
\mu(S) \leq1-\delta
.
\]
Then,
\[
\prob\bigl(\mu(S_n) \leq1\bigr) \rightarrow1
.
\]
\end{lemma}
\begin{pf}
By triangular inequality and thanks to the constraint on $\mu(S)$,
\[
\prob\bigl(\mu(S_n) \leq1\bigr) \geq\prob \bigl(\mu(S_n - S) \leq1 - \mu
(S) \bigr) \geq
\prob\bigl(\mu(S_n - S) \leq\delta\bigr)
,
\]
Convergence in probability of $S_n$
to $S$ concludes the proof:
\[
\prob\bigl(\mu(S_n - S) \leq\delta\bigr) \rightarrow1
\qquad\mbox{therefore }
\prob\bigl(\mu(S_n) \leq1\bigr) \rightarrow1
.
\]
\upqed
\end{pf}
Let us consider the full vector $\tildebbeta^n$ with coefficients
$\tildebbeta_\supp^n$ defined as in Lemma~\ref
{lemsupportconsistency} and
other coefficients null, $\tildebbeta_{\supp^c}^n=\mathbf{0}$.
We now proceed to the last step of the proof of
Theorem~\ref{thmsupportconsistency}, by proving that $\tildebbeta^n$ satisfies
the coop-Lasso optimality conditions with probability tending to 1 under
the additional conditions (A4)--(A5).
The final conclusion then results from the uniqueness of the coop-Lasso
estimator.
First, consider optimality conditions with respect to $\bbeta_{\supp
} $.
As a result of Lemma~\ref{lemsupportconsistency}, the probability that
$\tildebbeta_j^n \neq0$ for every $j \in\supp$ tends to 1.
Thereby, $\tilde{\bbeta}{}_{\supp}^n$ satisfies \eqref{eqoptimality}
on the restriction of $\mathbf{X}$ to covariates in $\supp$ with probability
tending to 1.
As $\tildebbeta_{\supp^c}^n=\mathbf{0}$,
then $\mathbf{X}\tildebbeta^n = \mathbf{X}_{\centerdot\supp
}\tildebbeta_{\supp}^n$
and for every $j \in\supp$,
$\|\bvarphi_j(\tildebbeta_{\supp_k})^n\| = \|\bvarphi_j(\tildebbeta
_{\group}^n)\|$,
therefore, $\tilde{\bbeta}{}_{\supp}^n$ satisfies \eqref{eqoptimality}
in the original problem with probability tending to 1.
Second, $\tildebbeta_{\supp^c}^n$ should also verify the optimality conditions
\eqref{eqoptimalityall} with probability tending to 1.
With assumption (A3), we only have to consider two cases that read:
\begin{itemize}
\item if group $k$ is excluded from the support, one must have
\begin{eqnarray}\label{eqoptcond2a}
&&\prob \bigl(\max \bigl( \bigl\|
\bigl((\mathbf{X}_{\centerdot\supp_k^c})^\intercal(\mathbf
{X}\tildebbeta^n-\mathbf{y})\bigr)^+
\bigr\|_n, \bigl\|
\bigl((\mathbf{X}_{\centerdot\supp_k^c})^\intercal(\mathbf
{X}\tildebbeta^n-\mathbf{y})\bigr)^-
\bigr\|_n \bigr) \leq\lambda_n w_k \bigr) \nonumber
\\[-8pt]
\\[-8pt]&& \qquad \rightarrow1
;
\nonumber\hspace*{-35pt}
\end{eqnarray}
\item if group $k$ intersects the support, with either positive
($\nu_k = 1$) or negative ($\nu_k=-1$) coefficients, one must have
\begin{equation}
\prob \bigl( \{\nu_k (\mathbf{X}_{\centerdot\supp_k^c})^\intercal
(\mathbf{X}\tildebbeta^n-\mathbf{y})
\succeq\mathbf{0}\} \cap\{ \|(\mathbf{X}_{\centerdot\supp
_k^c})^\intercal(\mathbf{X}\tildebbeta^n-\mathbf{y})\|_n
\leq\lambda_n w_k \} \bigr) \rightarrow1
.
\label{eqoptcond2b}\hspace*{-30pt}
\end{equation}
\end{itemize}
To prove \eqref{eqoptcond2a} and \eqref{eqoptcond2b}, we study the
asymptotics of
$(\mathbf{X}_{\centerdot\supp_k^c})^\intercal(\mathbf
{X}\tildebbeta^n-\mathbf{y})/n$
for any group such that $\supp_k^c$ is not empty. As a consequence of
the existence of the fourth order moments of the centered random
variables $X$ and\vadjust{\goodbreak}
$Y$, the multivariate central limit theorem applies, yielding
\begin{eqnarray*} \label{eqTCL}
\frac{\mathbf{X}^\intercal\mathbf{X}}{n} &=& \frac{1}{n}
\sum_{i=1}^n \mathbf{x}_{i}^\intercal\mathbf{x}_{i} =
\bPsi+ O_P(n^{-1/2}),\nonumber
\\[-8pt]
\\[-8pt]
\frac{\mathbf{X}^\intercal\boldsymbol{\varepsilon} }{n} &=& \frac{1}{n}
\sum_{i=1}^n \mathbf{x}_{i} \boldsymbol{\varepsilon}_i =
O_P(n^{-1/2})
\nonumber
\end{eqnarray*}
Then, we derive from \eqref{eqTCL} and the definition of $\tildebbeta
^n$ that
\begin{eqnarray}\label{eqkeyline1}
\frac{1}{n} (\mathbf{X}_{\centerdot\supp_k^c})^\intercal
(\mathbf{X}\tildebbeta^n-\mathbf{y}) & =&
\frac{1}{n} (\mathbf{X}_{\centerdot\supp_k^c})^\intercal\mathbf{X}
(\tildebbeta^n - \bbeta^\star) -
\frac{1}{n}(\mathbf{X}_{\centerdot\supp_k^c})^\intercal
\boldsymbol\varepsilon\nonumber\\
& =& \frac{1}{n}(\mathbf{X}_{\centerdot\supp_k^c})^\intercal
\mathbf{X}_{\centerdot\supp}
(\tildebbeta_{\supp}^n - \bbeta_{\supp}^\star) + O_P(n^{-1/2}) \\
& = &\bPsi_{\supp_k^c \supp}(\tildebbeta_{\supp}^n -
\bbeta_{\supp}^\star) + O_P(n^{-1/2})
,\nonumber
\end{eqnarray}
while the combination of \eqref{eqTCL} and optimality conditions
\eqref{eqoptimality} on $\tildebbeta_{\supp}^n$ leads to
\begin{equation}
\bPsi_{\supp\supp}(\tildebbeta_{\supp}^n-\bbeta_{\supp}^\star) =
- \lambda_n \mathbf{D}(\tildebbeta_\supp^n)\tildebbeta_\supp^n +
O_P(n^{-1/2})
,\label{eqkeyline2}
\end{equation}
where $\mathbf{D}(\cdot)$ is the weighting matrix \eqref{eqdefD}.
Put \eqref{eqkeyline1} and \eqref{eqkeyline2} together to finally obtain
\begin{equation}
\frac{1}{n}(\mathbf{X}_{\centerdot\supp_k^c})^\intercal(\mathbf
{X}\tildebbeta^n-\mathbf{y})
= - \lambda_n \bPsi_{\supp_k^c \supp}
\bPsi_{\supp\supp}^{-1}\mathbf{D}(\tildebbeta_\supp^n)\tildebbeta
_\supp^n +
O_P(n^{-1/2})\hspace*{-30pt}
. \label{eqOp}
\end{equation}
Now, define for any $k$ such that $\supp_k^c$ is not empty:
\[
R_{k,n} = \frac{1}{w_k \lambda_n}\frac{1}{n}
(\mathbf{X}_{\centerdot\supp_k^c})^\intercal(\mathbf{X}\tildebbeta
^n-\mathbf{y})
\quad\mbox{and} \quad R_k = - \frac{1}{w_k} \bPsi_{\supp_k^c \supp}
\bPsi_{\supp\supp}^{-1}\mathbf{D}(\bbeta^\star_\supp)\bbeta
^\star_\supp
.
\]
Limits \eqref{eqoptcond2a} and \eqref{eqoptcond2b} are expressed:
\begin{itemize}
\item if group $k$ is excluded from the support, one must have
\[
\prob \bigl(\max (\|R_{k,n}^+\|,\|R_{k,n}^-\| )\leq
1 \bigr) \rightarrow1
;
\]
\item if group $k$ intersects the support, with either positive
($\nu_k = 1$) or negative ($\nu_k=-1$) coefficients, one must have
\[
\prob \bigl( \{ \nu_kR_{k,n} \succeq\mathbf{0}\} \cap\{\|(\nu
_kR_{k,n})^+\| \leq1\} \bigr)
\rightarrow1
.
\]
\end{itemize}
Remark that, as a continuous function of $\tildebbeta_\supp^n$,
$\mathbf{D}(\tildebbeta_\supp^n)\tildebbeta_\supp^n$ converges in
probability to
$\mathbf{D}(\bbeta^\star_\supp)\bbeta^\star_\supp$. Therefore,
with a decrease rate
for $\lambda_n$ chosen such that $n^{1/2}\lambda_n
\rightarrow\infty$, equation \eqref{eqOp} implies
\begin{equation}
R_{k,n} \inprob R_k . \label{eqconvproba}
\end{equation}
It now suffices to successively apply Lemma \ref
{lemUseIrrepresentableCond} to
the appropriate vectors and norms to show that $\tildebbeta_{\supp
^c}^n$ satisfies
\eqref{eqoptcond2a} and \eqref{eqoptcond2b}:
\begin{itemize}
\item if group $k$ is excluded from the support, (A4) assumes that
there exists
$\eta>0$, such that
\[
\max(\|R_k^+\|,\|R_k^-\|) \leq1 -\eta\vadjust{\goodbreak}
,
\]
and Lemma \ref{lemUseIrrepresentableCond} applied to $\mu(u) =
\max(\|u^+\|,\|u^-\|)$ provides
\[
\prob\{\max(\|R_{k,n}^+\|,\|R_{k,n}^-\|) \leq1\}
\rightarrow1
.
\]
\item if group $k$ intersects the support, with either positive
($\nu_k = 1$) or negative ($\nu_k=-1$) coefficients,
\begin{eqnarray*}
&& \prob \bigl( \{\|(\nu_kR_{k,n})^+\| \leq1\} \cap\{ \nu_kR_{k,n}
\succeq\mathbf{0}\} \bigr)
\\
& & \qquad = 1 - \prob
\bigl(\{\|(\nu_kR_{k,n})^+\| > 1 \} \cup\{ \nu_kR_{k,n} \prec
\mathbf{0} \} \bigr) \\
&& \qquad \geq1 - \prob
\bigl (\|(\nu_kR_{k,n})^+\| > 1 \bigr) - \prob ( \nu_kR_{k,n}
\prec\mathbf{0} ) \\
&& \qquad \geq1 - \prob \bigl( \max(\| R_{k,n}^+\|,\|
R_{k,n}^-\|) > 1
\bigr) - \prob ( \nu_kR_{k,n} \prec\mathbf{0} )
.
\end{eqnarray*}
As previously, the first probability in the sum tends to 0 because of
(A4) and Lemma
\ref{lemUseIrrepresentableCond}. The second
probability tends to 0 from (A5) and of the convergence in probability
of $R_{k,n}$ to $R_k$. Therefore, the overall probability tends to 1.
\end{itemize}
Denote by $A_{k,n}$ these events on which coefficients in $\supp_k^c$
are set to 0. We just showed that individually for each group $k$
with true null coefficients, $P(A_{k,n}) \rightarrow1$. This implies that
\[
\prob \biggl( \bigcup_{k\dvtx \supp_k^c \neq\varnothing} A^c_{k,n} \biggr)
\leq\sum_{k\dvtx \supp_k^c \neq\varnothing} \prob ( A_{k,n}^c
)
\rightarrow0,
\]
which in turn concludes the proof:
\[
\prob \biggl(\bigcap_{k\dvtx \supp_k^c \neq\varnothing} A_{k,n} \biggr)
\rightarrow1.
\]
\end{appendix}
\section*{Acknowledgments}
We would like to thank Marine Jeanmougin for her helpful comments on
the breast cancer data set and for sharing her differential analysis
on the subset of basal tumors. We also thank Catherine Matias for her
careful reading of the manuscript and Christophe Ambroise for fruitful
discussions.
|
\section{Introduction}
Perovskite oxide is denoted as $AB$O$_3$, where $A$ and $B$ represent lanthanide (and/or alkaline earth
elements) and transition-metal elements, respectively. $B$ ion is surrounded by six oxygen ions, and
$B$O$_6$ octahedron is formed. The octahedron mainly contributes electrical and magnetic proerties
of the perovskite oxide. According to the ionic radius of $A$ ion, the bond angle $\angle$ $B$-O-$B$
deviates from 180$^{\circ}$, which causes a change of the bandwidth.
In a substituted system of $R^{3+}_{1-x}$$A^{2+}_{x}$$B$O$_3$, the carrier concentration
is also controlled as well as the bandwidth.
The A-site ordered manganese oxide $R$BaMn$_2$O$_6$ ($R:$ lanthanide)
has attracted much attention because of significant physical phenomena
such as the large magnetoresistance of 1,000 \% at room temperature \cite{nakajima0},
charge and orbital orderings at high temperatures \cite{nakajima1}.
These properties attribute to
A-site ordering working as "a periodic" Coulomb potential
which stabilizes charge, spin, and/or orbital orderings of the electrons on B-site \cite{motome}.
Owing to the randomness, A-site disordered phase of $R_{0.5}$Ba$_{0.5}$MnO$_3$ displays
the spin-glass state or the itinerant ferromagnetic state instead of the charge ordered state seen in the ordered phase
\cite{nakajima1,akahoshi}.
The A-site ordered cobalt oxide Sr$_3$YCo$_4$O$_{10.5}$ also exhibits peculiar
magnetic properties with a high ferromagnetic (ferrimagnetic) transition
temperature of 335 K \cite{kobayashi2}, which is in contrast with a spin state crossover
near 100 K in LaCoO$_3$ \cite{asai}.
In this compound, the A-site ordering stabilizes oxygen deficient ordering,
which makes a volume of the CoO$_6$ octahedron larger than that of LaCoO$_3$ \cite{ishiwata} and
gives a different coordination number of Co$^{3+}$ from LaCoO$_3$. These modifications
stabilize high-spin and/or intermediate-spin states of Co$^{3+}$ even
at low temperatures, which causes the peculiar magnetic properties.
Hence, partial substitution for the A-site or B-site cations strongly affects the spin state leading to significant suppression
of the magnetic order; Ca substitution for Sr site \cite{yoshida}
and 6\%-Mn doping in the B site \cite{kobayashi4} destroys the room-temperature ferromagnetism of
Sr$_3$YCo$_4$O$_{10.5}$.
Recently, a new A-site ordered perovskite Y$_{0.8}$Sr$_{2.2}$Mn$_2$GaO$_{7.9}$
(Sr$_{2.93}$Y$_{1.07}$ Mn$_{2.66}$Ga$_{1.34}$O$_{10.53}$) was reported
by Gillie {\it et al.} \cite{gillie}, which is isostructural to Sr$_{3}$YCo$_4$O$_{10.5}$ \cite{istomin1,withers}.
As shown in Fig. 1, this compound has an octahedral site where Mn$^{3+}$ ions mainly occupy
with about 10\%-Ga ions intermixed and a tetrahedral site where both Mn$^{3+}$ and Ga$^{3+}$ ions occupy.
They found an antiferromagnetic state below 100 K in this material showing
that Y$_{0.8}$Sr$_{2.2}$Mn$_2$GaO$_{7.9}$ was a antiferromagnetic insulator,
however, they did not report the transport properties.
We have prepared polycrystalline samples of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=$ 0, 0.5, 1, 2)
where $d$ represents oxygen deficiency
and investigated the transport and magnetic properties in relation to the strcture.
We have succeeded in preparing both A-site ordered and disordered phases for $x=1$ using
different preparation conditions and observe a significant decrease of the resistivity in the disordered phase.
We attribute this to a change of dimensionality in conduction and/or A-site disordered effect.
This material can be a good playground to study order-disorder effect on the electronic states of the
antiferromagnetic insulator with Mn$^{3+}$.
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=4cm,clip]{fig1.eps}
\caption{(Color online) Crystal structure of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x\geq 1$).
}
\end{center}
\end{figure}
\section{Experiments}
Polycrystalline samples of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2)
were prepared by a solid state reaction. Stoichiometric amounts of
SrCO$_3$, Er$_2$O$_3$, Mn$_3$O$_4$, and Ga$_2$O$_3$ were mixed,
and the mixture was sintered at 1250 $^{\circ}$C for 12 h for $x=0$ and 0.5, 6 h for $x=1$ and 2
in N$_2$ flow ($100-200$ ml/min).
Then, the product was finely ground, pressed into a pellet, and sintered at 1250 $^{\circ}$C for 12 h for $x=0$,
24 h for $x=0.5$, and 10 h for $x=1$ and 2 in N$_2$ flow ($100-200$ ml/min).
The second process was repeated 2 times for $x=0.5$, and once for $x=1$ and 2
with intermediate grindings and pelletizings.
As shown in Fig. 2(c), $x=1$ sample exhibits A-site ordered structure. We sintered the $x=1$ sample
again using the second process, and found the sample shows a disordered structure shown in Fig. 3.
The x-ray diffraction of the sample was measured using a standard diffractometer with Cu K$\alpha$
radiation as an x-ray source in the $\theta -2\theta $ scan mode.
The structural simulations were performed using a RIETAN-2000 program \cite{izumi}.
The resistivity was measured by a four-probe method in a liquid He cryostat.
The thermopower was measured using a steady-state technique in a liquid He cryostat
with copper-constantan thermocouple to detect a small temperature
gradient of about 1 K/cm. The magnetization was measured from 5 to 400 K by a
commercial superconducting quantum interference device (SQUID, Quantum Design MPMS).
\section{Results and discussion}
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=6cm,clip]{fig2.eps}
\caption{(Color online) (a)-(d) X-ray diffraction patterns of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2).
The insets represent the magnified x-ray patterns at low angles.
The red line of the inset of Fig. 2(d) represents a simulated pattern.
}
\end{center}
\end{figure}
Figure 2 shows the x-ray diffraction patterns of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ ($x=0, 0.5, 1$, and 2).
All the peaks are indexed as a cubic cell of the space group $Pm3m$ with the lattice parameter of
$a\sim $ 3.8 and 3.85 \AA~for $x=$ 0 and 0.5, respectively. This cubic cell is also seen in
Sr$_{1-x}$Y$_{x}$CoO$_{3-\delta }$ with small $x$ \cite{kobayashi2}.
With increasing Ga content $x$, crystal structure changes from the cubic perovskite
to a tetragonal A-site ordered perovskite (space group: $I4/mmm$, Fig. 1) with the lattice parameter of
$a\sim $ 7.63 and 7.65 \AA,~and $c\sim $ 15.58 and 15.56 \AA~for $x=$ 1 and 2, respectively.
This result is consistent with the structural analysis by Gillie {\it et al.} \cite{gillie}.
As shown in the inset of Fig. 2(d), superstructure peaks corresponding to the A-site ordering are observed,
while they do not appear for $x=$ 0 and 0.5 samples.
Ga ions selectively occupy the tetrahedral site to stabilize the A-site ordered structure as shown in Fig. 1 \cite{gillie}.
Thus, $x=1$ is a minimal amount to stabilize the structure.
Gillie {\it et al.} reported that the oxygen content of Sr$_{2}$YMn$_{2}$GaO$_{7.9}$ is
7.9 showing formal valence of Mn ion is almost 3+.
Thus, it is assumed that the formal valence of Mn ion is also 3+ in the ordered compounds presented here.
Figure 3 shows the x-ray diffraction patterns of Sr$_{3}$ErMn$_{3}$GaO$_{10.5-d}$
with different preparation conditions. Obviously, two patterns are different;
one sample was identified to the tetragonal ordered phase (hereafter this is denoted by O sample),
and the other was identified to the disordered-cubic perovskite phase (D sample).
Similar structures are originally found in
$R$BaMn$_{2}$O$_{6}$/$R_{0.5}$Ba$_{0.5}$MnO$_3$ systems \cite{nakajima1}.
We would like to say that the O sample is metastable so that the longer sintering stabilizes the disordered phase.
Thus, this composition is just on the verge of order and disorder.
As shown in the inset of Fig. 3, the superstructure peaks of D sample are hardly visible.
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=6.5cm,clip]{fig3.eps}
\caption{(Color online) X-ray diffraction patterns of ordered and disordered Sr$_{3}$ErMn$_{2}$Ga$_2$O$_{10.5-d}$.
The inset represents the magnified x-ray pattern at low angles.
}
\end{center}
\end{figure}
Figure 4 (a) shows the thermopower of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
($x=0, 0.5, 1$, and 2). The magnitude of the thermopower is
between $60$ and $110$ $\mu $V/K, and the sign is negative showing that the carriers are electrons.
Assuming that the formal Mn valence is 3+, we expect that a tiny amount of electrons on
Mn$^{2+}$ moves in the background of Mn$^{3+}$ as
shown in the inset of Fig. 4(a).
Using an extended Heikes formula \cite{koshibae},
the valence of Mn ion was evaluated to be 2.71+ at 300 K corresponding to $d=0.43$
for $x=1$ sample with spin
degeneracy term of $g_{\rm Mn^{2+}}=$6 and $g_{\rm Mn^{3+}}=$5.
Figure 4 (b) shows the resistivity of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
($x=0, 0.5, 1$, and 2). Semiconducting temperature dependence
is observed for all the samples. With $x$, the magnitude of the resistivity decreases mainly due to
decrease of scattering centers of the Ga ions. As seen in the inset of Fig. 4(b), the temperature
dependence is described by an activation-type conduction
$\rho =\rho _0$exp($\frac{E_{\rm g}}{k_{\rm B}T}$) where $E_{\rm g}$ represents activation energy above 200 K.
The activation energy $E_{\rm g}$ was evaluated to be 0.133, 0.146, 0.149, and 0.246 eV for
$x=$0, 0.5, 1, and 2, respectively.
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=6cm,clip]{fig4.eps}
\caption{(Color online) (a) Thermopower and (b) resistivity of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
($x=$0, 0.5, 1, and 2).
}
\end{center}
\end{figure}
Figure 5(a) shows the magnetization of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
($x=0, 0.5, 1$, and 2). As shown in Figs. 5(b) and (c), the data was fitted by the Curie-Weiss law described by
$\chi =\frac{C}{T-\theta _{\rm W}}+\chi _0$, where $C$, $\theta _{\rm W}$, and $\chi _0$
represent Curie constant, Weiss temperature and temperature independent term of the magnetic susceptibility,
respectively. $C$ was evaluated to be 0.023, 0.019, 0.026, and 0.018 emu/K$\cdot $g for $x=0, 0.5, 1$, and 2,
corresponding to 12.34, 11.14, 13.17, and 11.07 $\mu _{\rm B}$/f.u., respectively.
These values are roughly explained by coexistence of 9.6 $\mu _{\rm B}$
of Er$^{3+}$ with $g=\frac{6}{5}$ and $J=\frac{15}{2}$ and 3.87 $\mu _{\rm B}$ and 4.90 $\mu _{\rm B}$
in the high-spin states of Mn$^{4+}$ and Mn$^{3+}$.
$\chi _0$ was evaluated to be 1.39$\times$10$^{-5}$, 1.57$\times$10$^{-5}$, 0, and 0 emu/g for
$x=0, 0.5, 1$, and 2 samples, respectively. Since $x=0$ and 0.5 samples show relatively good
electric conductance compared with those of $x=1$ and 2 samples,
this contribution may come from conducting electrons.
All the samples exhibit negative $\theta _{\rm W}$
($-$45, $-$29, $-$52, and $-$30 K for $x=0, 0.5, 1$, and 2, respectively)
implying antiferromagnetic interaction in this system.
At around 60 K, the slope of $\chi ^{-1}$ changes showing an existence of magnetic anomaly for all the samples.
Since the antiferromagnetism was observed in Sr$_2$YMn$_2$GaO$_{8-d}$ below 100 K \cite{gillie},
the anomaly at around 60 K in Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
can be also related to antiferromagnetic order of Mn$^{3+}$.
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=6cm,clip]{fig5.eps}
\caption{(Color online) Magnetic susceptibility of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$
($x=$0, 0.5, 1, and 2). The inset shows inverse susceptibility.
}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\vspace*{0cm}
\includegraphics[width=7cm,clip]{fig6.eps}
\caption{(Color online) (a) Resistivity and (b) thermopower and magnetization of the ordered and disordered phases.
}
\end{center}
\end{figure}
Lastly, we will discuss a difference of the transport and magnetic properties between
the ordered (O sample) and disordered samples (D sample).
Figure 6 (a) shows the temperature dependences of the resistivity of the O and D samples.
The magnitude of the resistivity of the D sample is one order of magnitude smaller than that of the O sample,
while the thermopower and magnetization of the D sample quite resemble those
of the O sample as seen in Fig. 6(b).
This strongly contrasts with the difference between the ferromagnetic-metal state of
the disordered $R_{0.5}$Ba$_{0.5}$MnO$_3$ and charge-ordered
insulating state of the ordered $R$BaMn$_2$O$_6$ \cite{nakajima1}.
This variety of the properties comes from Mn$^{3.5+}$ with the charge degree of freedom,
while Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ has
Mn$^{3+}$ without the charge degree of freedom causing the small difference between the properties.
The difference of the resistivity shown in Fig. 6(a) is explained by several possible origins as follows:
(1) a decrease of the carrier concentration, (2) an increase of the scattering time,
(3) a change of dimensionality in conduction and (4) a change of electronic structure induced by the disordering of A site.
As seen in Fig. 6(b), two samples exhibit almost the same magnitude of the thermopower of $-100$ $\mu$V/K.
Since thermopower is a function of carrier concentration, the result shows
that carrier concentration does not differ so much in the two samples.
In addition, since the content $x$ of Ga ions which can be scattering centers is 1 in both samples,
a possibility of (2) can be also denied.
As stated in the introduction, Ga ion selectively occupies tetrahedral sites in the O phase, while
Ga and Mn ions randomly occupies in the D phase. Thus, a dimensionality in conduction can change from 2D in the
O sample to 3D in the D sample, which can be a possible origin of the decrease of the resistivity.
Another possibility is a change of electronic structure induced by the disordering of A site.
A-site disordering generally induces a random potential in a perovskite oxide, which
may affect the electronic structure of the material.
Indeed, it is theoretically found that the introduced random potential in the Hubbard Hamiltonian
causes an antiferromagnetic metallic phase instead of the antiferromagnetic insulating state \cite{shinaoka}.
Although the origin of the difference of the resistivity is not clear at present,
this system can be a good playground for investigating A-site disorder effect on antiferromagnetic
insulator with Mn$^{3+}$.
\section{summary}
In summary, we have measured x-ray diffraction, resistivity, thermopower, and magnetization
of Sr$_{3}$ErMn$_{4-x}$Ga$_x$O$_{10.5-d}$ system,
in which A-site ordered tetragonal phase appears above $x=1$, and observed
large negative thermopower, semiconducting conduction, and magnetic susceptibility with a kink at
60 K implying antiferromagnetism.
We succeed to prepare both A-site ordered and disordered phases for $x=1$ sample
and observe a significant decrease of the resistivity in the disordered phase.
We attribute this to a change of dimensionality in conduction and/or a change of electronic structure induced
by the disordering of A sites.
\section{acknowledgements}
We acknowledge I. Terasaki for fruitful discussion. This study was supported by the program entitled "Promotion of
Environmental Improvement for Independence of Young Researchers" under the
Special Coordination Funds for Promoting Science and Technology provided by
MEXT, Japan.
|
\section{Introduction}\label{s.introduction}
In the late sixties, Newhouse constructed the first examples of
$C^2$-open sets of non-hyperbolic surface diffeomorphisms. Any such set
$\cU$ consists of diffeomorphisms with \emph{$C^2$-robust
homoclinic tangencies}:
every diffeomorphism $f\in \cU$ has a hyperbolic set
$K_f$ (depending continuously on $f$) whose stable and unstable
manifolds have non-transverse intersections, see \cite{N68}.
Later, in \cite{N79}, Newhouse proved that homoclinic tangencies of
surface diffeomorphisms can be {\it{stabilized\/}}:
given a diffeomorphism $f$ with
a homoclinic tangency associated to a saddle $p_f$,
there is a $C^2$-open set whose closure contains $f$
and which consists of diffeomorphisms $g$ with robust homoclinic tangencies
associated to hyperbolic sets
$K_g$ containing the continuation $p_g$ of $p_f$.
In particular, these results show that homoclinic tangencies always generate
$C^2$-robust homoclinic tangencies. In fact, robust homoclinic
tangencies are present in all known examples of open sets of
non-hyperbolic surface diffeomorphisms. Let us observe that
homoclinic tangencies of $C^1$-diffeomorphisms
defined on surfaces cannot be stabilized, see \cite{Gugu}.
Similarly, all known examples of $C^1$-open sets formed by
non-hyperbolic diffeomorphisms
exhibit $C^1$-robust {\emph{heterodimensional cycles,}} that is,
cycles relating the invariant manifolds of two hyperbolic sets of
different $s$-indices (dimension of the stable bundle). Note that
the existence of such cycles can only occur in dimension $\ge 3$.
We wonder if, as in the case of homoclinic tangencies of
$C^2$-diffeomorphisms, heterodimensional cycles can be made
$C^1$-robust and can be $C^1$-stabilized. A first partial answer to this
question is given in \cite{BD08}: heterodimensional cycles associated to
periodic saddles whose indices differ by one
generate (by arbitrarily small $C^1$-perturbations)
$C^1$-robust heterodimensional cycles.
In some extend, the results in \cite{BD08} are a version of the ones
by Newhouse in \cite{N68,N79} in the context of $C^1$-heterodimensional
cycles.
However, compared with Newhouse's results for homoclinic tangencies, the ones in
\cite{BD08} have an important disadvantage: While the hyperbolic
sets with the robust homoclinic tangencies in \cite{N68,N79}
contain continuation of the saddle with the initial tangency,
the hyperbolic sets involved in robust cycles in \cite{BD08} do in general not contain the continuations of the saddles in the initial cycle. However this precise question can be important
for understanding the global dynamics of non-hyperbolic
diffeomorphisms.
Let us discuss more this question in more detail.
Following Conley theory \cite{Con}
and motivated by spectral decomposition theorems \cite{Newhouse,A03},
this global dynamics is structured using
{\emph{homoclinic}} or/and \emph{chain recurrence classes} as
``elementary" pieces of dynamics, see the
definitions below.
One aims
to describe the dynamics of each piece and the relations between
different pieces (cycles), for further details see \cite[Chapter
10.3-4]{BDVbook} and \cite{bible}.
In general,
the homoclinic class of a hyperbolic periodic
point is contained in its chain recurrence class.
An important property is that for $C^1$-generic diffeomorphisms
homoclinic classes
and chain recurrence classes of periodic points coincide,
\cite[Remarque 1.10]{BC04}. However,
in non-generic situations, two different homoclinic classes (even of saddles of different indices) may be joined by a cycle,
hence they are contained in the same chain recurrence class.
A question is when one can join them in a $C^1$-robust way by small
perturbations. This occurs if the cycle can be
stabilized. For instance, this is specially important for understanding the indices of the periodic points in an elementary piece of dynamics.
While the above explains why the stabilization of cycles is relevant,
let us now provide the precise
definitions of the concepts involved. First, recall that a hyperbolic basic set $K$ of a
diffeomorphism $f$ has a (uniquely defined) {\emph{continuation}}
$K_g$ for all $g$ close to $f$: $K_g$ is a hyperbolic basic set, close to $K$, and the dynamics of $f|_{K}$ and $g|_{K_g}$ are conjugate. The {\emph{$s$-index}} of a hyperbolic transitive set is the dimension of its stable bundle.
\begin{defi}[Robust continuations of cycles]
{\em{$\,$
{$\bullet$} The diffeomorphism $f$ is said to have a {\emph{heterodimensional
cycle}} associated to hyperbolic basic sets
$K$ and $L$ if these sets have different $s$-indices and their
stable and unstable manifolds meet cyclically:
$$
W^s(K,f)\cap W^u(L,f)\ne\emptyset \quad \mbox{and} \quad
W^u(K,f)\cap W^s(L,f)\ne\emptyset.
$$
{$\bullet$} The cycle associated to $K$ and $L$ is {\emph{
$C^1$-robust}} if there is a $C^1$-neighborhood $\cU$ of $f$ such
that for all $g\in \cU$ the hyperbolic continuations $K_g$ and
$L_g$ of $K$ and $L$ have a heterodimensional cycle.
{$\bullet$} A heterodimensional cycle associated to a pair of
saddles $p$ and $q$ of $f$ {\emph{can be $C^1$-stabilized}} if
every $C^1$-neighborhood $\cU$ of $f$ contains a diffeomorphism
$g$ with hyperbolic basic sets $K_g\ni p_g$ and $L_g\ni q_g$ having a robust heterodimensional cycle. Here $p_g$ and $q_g$ are the continuations of
$p$ and $q$ for $g$.}}
\end{defi}
\begin{defi}[Fragile cycle]{\em{
A heterodimensional cycle associated to a pair of saddles is
{\emph{$C^1$-fragile}} if it cannot be $C^1$-stabilized.}}
\end{defi}
The previous discussion leads to the following
question that we address in this paper:
{\emph{
Can every heterodimensional cycle be
$C^1$-stabilized?}}
\smallskip
The results in \cite{BDK}
give a positive answer to this question for ``most" types of
{\emph{coindex one}}
heterodimensional cycles, that is, related to saddles whose $s$-indices differ by one. Indeed in \cite{BDK} it is proved that fragile coindex one cycles associated to saddles $p$ and $q$ exhibit a quite specific geometry:
\begin{itemize}
\item
The homoclinic classes of $p$ and $q$ are both trivial.
\item
The {\emph{central eigenvalues}} of $p$ and $q$ are all real and
positive.\footnote{The definition of central eigenvalues is a little
intricate. Assuming that the $s$-index of $p$ is bigger than the
one of $q$, the central eigenvalues correspond to the weakest
contracting direction of $p$ and the weakest expanding direction
of $q$.}
\item
There is a well-defined one-dimensional orientable central bundle
$E^c$ along the cycle (i.e. defined on some closed set containing
the saddles $p$ and $q$ in the cycle and a pair of heteroclinic
orbits $x\in W^s(p)\cap W^u(q)$ and
$y\in W^u(p)\cap W^s(q)$), but the cycle diffeomorphism does not preserve
the orientation of $E^c$. The cycle is \emph{twisted} by the terminology in
\cite{ASY06}.
\end{itemize}
In this paper we provide examples of fragile coindex one cycles, see
Theorem~\ref{t.fragil}.
\subsection{Definitions and statement of results}
Recall that the {\emph{homoclinic class}} of a hyperbolic periodic
point $p$, denoted by $H(p,f)$, is the closure of the transverse
intersections of the stable and unstable manifolds of the orbit of
$p$. The homoclinic class $H(p,f)$ coincides with the closure of the set of all saddles
$q$ {\emph{homoclinically related with $p$,}} i.e. the stable
manifold of the orbit of $q$ transversely meets the unstable
manifold of the orbit of $p$ and vice-versa. A homoclinic class is
{\emph{non-trivial}} if it contains at least two different orbits.
Let us now recall the definition of a {\emph{chain recurrence class.}}
A finite sequence of points $(x_i)_{i=0}^n$ is an
{\emph{$\epsilon$-pseudo-orbit}} of a diffeomorphism $f$ if
$\mbox{dist}(f(x_i),x_{i+1})<\epsilon$ for all $i=0,\dots,n-1$. A
point $x$ is {\emph{chain recurrent}} for $f$ if
for every $\epsilon>0$ there is an $\epsilon$-pseudo-orbit
$(x_i)_{i=0}^n$, $n\ge 1$,
starting and ending at $x$ (i.e. $x=x_0=x_n$). The {\emph{chain recurrent set}} $R(f)$ of $f$ is the set of all chain recurrent points. This
set splits into disjoint {\emph{chain recurrence classes}}: the
class $C(x,f)$ of $x\in R(f)$ is the set of points $y$ such that
for every $\epsilon>0$ there are $\epsilon$-pseudo-orbits joining
$x$ to $y$ and $y$ to $x$.
A periodic point $p$ of $f$ is {\emph{isolated}} if its {\emph{chain
recurrence class}} coincides with its orbits. In this case, the
orbit of $p$ is the maximal invariant set in some \emph{filtrating
neighborhood}. This implies that the homoclinic class of $p$
is $C^1$-robustly trivial (i.e. the homoclinic class of $p_g$ is
trivial for every $g$ close to $f$).
We are now ready to state our main result.
\begin{theo}\label{t.fragil}
There is an open set $\cU$ of ${\operatorname{Diff}}^1 (\SS^2\times \SS^1)$ and a
codimension one submanifold $\Sigma$ contained in $\cU$ with the
following property: For every $f\in \cU$ there are hyperbolic
saddles $p_f$ and $q_f$ with
different $s$-indices depending continuously on $f$ such that
\begin{enumerate}
\item every $f\in \Sigma$ has
a heterodimensional cycle associated to $p_f$ and $q_f$,
\item the set $\cU\setminus\Sigma$ is the union of two connected sets $\, \cU^+$ and
$\,\cU^-$ such that
\begin{itemize}
\item
for every $f\in \cU^+$ the saddle $p_f$ is isolated,
\item
for every $f\in \cU^-$ the saddle $q_f$ is isolated.
\end{itemize}
\end{enumerate}
\end{theo}
Note that if two hyperbolic basic sets $K_f$ and $L_f$ have a
heterodimensional cycle then the chain recurrence classes of any
pair of saddles $a_f\in K_f$ and $b_f\in L_f$ coincide. In
particular, if the cycle associated to $K_f$ and $L_f$ is robust
then the chain recurrence classes of $a_g$ and $b_g$ are the same
for all $g$ in some neighborhood of $f$. In particular, the chain recurrence
class $C(a_g,g)=C(b_g,g)$ is non-trivial and the saddles $a_g$ and
$b_g$ are both non-isolated for $g$. Thus
Theorem~\ref{t.fragil} implies that the cycles in $\Sigma$ cannot be made
robust. This implies the following:
\begin{Coro}
\label{c.fragil} The submanifold $\Sigma$ in Theorem~\ref{t.fragil}
consists of diffeomorphisms having $C^1$-fragile
heterodimensional cycles.
\end{Coro}
As for any $n>3$ the set $\SS^2\times \SS^1$ can be embedded as a normally contracting manifold in a ball $\BB^n$, we obtain the
following.
\begin{Coro}
Any compact manifold $M$ with $\dim M>3$ supports diffeomorphisms
with $C^1$-fragile cycles.
\end{Coro}
As our examples demand a somewhat specific topological
configuration, the following question arises naturally.
\begin{Ques} Does every $3$-manifold admit diffeomorphisms with $C^1$-fragile
heterodimensional cycles?
\end{Ques}
The examples presented in this paper display many interesting and somehow unexpected properties. There are also many important aspects of their dynamics yet unexplored. Thus, after completing our construction, in Section~\ref{s.conclusion} we conclude with a discussion about the properties of our
examples.
\medskip
This paper is organized as follows.
In the first step of our construction, in Section~\ref{s.auxiliary},
we build an auxiliary Morse-Smale
vector field $X$ on the $3$-sphere $\SS^3$.
In Section~\ref{s.surgery}, we consider a surgery
in $\SS^3$
(associated to some identifications by
a local diffeomorphism $\Psi$ of $\SS^3$). This surgery provides
a diffeomorphism $F_\Psi$ defined on $\SS^2 \times \SS^1$
induced by the time-one map $F_0=X_1$ of the vector field $X$
and the gluing map $\Psi$. We also see how the dynamics of $F_\Psi$
depends on the gluing map $\Psi$. In Section~\ref{ss.neighborhood}, we study the dynamics of
diffeomorphisms close to $F_\Psi$. Finally, in Section~\ref{s.psi}, we
choose the gluing map $\Psi$ to get a diffeomorphism $F_\Psi$ with a fragile cycle
and construct the submanifold $\Sigma$ consisting of diffeomorphisms with
fragile cycles. The paper is closed with a discussion section.
\section{An auxiliary vector field on
$\SS^3$}\label{s.auxiliary}
In this section we construct a Morse-Smale vector field
defined on the three sphere $\SS^3$
whose non-wandering set consists of
singular points.
This vector field also
satisfies some normally hyperbolic properties. We now go to the
details of this construction.
We consider the sphere $\SS^3$ as the union of
two solid tori $\cT_1$ and $\cT_2$ with the same boundary
$\partial \cT_1=\partial \cT_2={\mathbb T}^2$. A simple closed curve of
${\mathbb T}^2$ is a \emph{$\cT_i$-meridian} if it is not $0$-homotopic in
${\mathbb T}^2$ but is $0$-homotopic in $\cT_i$.
We consider an identification of the boundaries of these solid tori
that does not preserve the meridians: the
$\cT_2$-meridians are isotopic in $\cT_1$ to the ``central circle"
of $\cT_1$ and are classically called \emph{$\cT_1$-parallels}.
Similarly, $\cT_1$-meridians are $\cT_2$-parallels.
\subsection{An auxiliary Morse-Smale vector field $X$ in $\SS^3$}\label{ss.auxiliary}
Consider a Morse-Smale vector field $X$ defined on $\SS^3$ such
that (see Figure~\ref{f.MorseSmale}):
\begin{figure}[htb]
\psfrag{s1}{$s_1$}
\psfrag{s2}{$s_2$}
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{r1}{$r_1$}
\psfrag{r2}{$r_2$}
\psfrag{q1}{$q_1$}
\psfrag{q2}{$q_2$}
\psfrag{su}{$\sigma^u$}
\psfrag{ss}{$\sigma^s$}
\psfrag{x}{$X$}
\psfrag{g1s}{$\gamma_1^s$}
\psfrag{g2s}{$\gamma_2^s$}
\psfrag{T2}{$\cT_2$}
\psfrag{T1}{$\cT_1$}
\psfrag{T}{${\mathbb T}$}
\includegraphics[width=7.5cm]{morsesnalerev.eps}\hspace{1cm}
\label{f.MorseSmale}
\caption{The Morse-Smale vector field $X$ defined on $\SS^3$.}
\end{figure}
\begin{enumerate}
\item
$X$ is transverse to $\partial \cT_1=\partial \cT_2={\mathbb T}^2$.
\item
The solid torus $\cT_1$ is attracting and the solid torus $\cT_2$
is repelling: The positive orbit of any point $x\in {\mathbb T}^2$ enters
(and remains) in the interior of $\cT_1$ and its negative orbits
enters (and remains) in the interior of $\cT_2$.
\item
\label{i.circles}
The maximal invariant set of $X$ in $\cT_1$ is a normally
hyperbolic (contracting) circle $\sigma^s$. Analogously, the
maximal invariant set of $X$ in $\cT_2$ is a normally hyperbolic
(repelling) circle $\sigma^u$.
\item The limit set of $X$ is
$\{s_1,s_2,r_1,r_2,p_1,p_2,q_1,q_2\}$, where
\begin{itemize}
\item
$s_1,s_2\in \sigma^s$ are attracting singularities,
\item
$r_1,r_2\in \sigma^u$ are repelling singularities,
\item
$p_1,p_2\in \sigma^s$ are saddle singularities of $s$-index $2$,
and
\item
$q_1,q_2\in \sigma^u$ are saddle singularities of $s$-index $1$.
\end{itemize}
\item
\label{i.Xsepartices} The two (one-dimensional) separatrices
of the unstable manifold of the singularity $p_i$,
$i=1,2$, are contained in
$W^s(s_1)$ and in $W^u(s_2)$. A similar
assertion holds for the separatrices of the stable manifold of
$q_i$, $i=1,2$, that are contained in $W^u(r_1)$ and $W^u(r_2)$.
\item The local stable manifold of $p_i$, $i=1,2$, is a
$2$-disk
contained in $\cT_1$ whose boundary
\[
\partial W^s_{loc}(p_i) =
W^s_{loc}(p_i)\cap
\partial\cT_1\stackrel{\scriptscriptstyle\rm def}{=} \gamma_i^s
\]
is a $\cT_1$-meridian. Similarly,
the local unstable manifold of $q_i$, $i=1,2$, is a $2$-disk
contained in $\cT_2$ whose boundary
\[
\partial
W^u_{loc}(q_i)=W^u_{loc}(q_i)\cap
\partial \cT_2\stackrel{\scriptscriptstyle\rm def}{=} \gamma_i^u,
\]
is a $\cT_2$-meridian.
\item
\label{i.points} For every $i,j\in\{1,2\}$, the curve $\gamma_i^s$
is transverse to $\gamma_j^u$ and the intersection $\gamma_i^s\cap
\gamma_j^u$ is exactly one point $x_{ij}$.
\end{enumerate}
\begin{rema}
[Dynamics of the vector field $X$] \label{r.stablecylinders} $\,$
\em{
\begin{enumerate}
\item \label{i.boundary}
The boundary ${\mathbb T}^2$ of the solid torus $\cT_1$ is the union of
two cylinders $\mbox{l\hspace{ -.47em}C}_1^s$ and $\mbox{l\hspace{ -.47em}C}_2^s$ with disjoint interiors and
the same boundary $\gamma_1^s \cup \gamma_2^s$. The notation is chosen
such that $W^s(s_i)\cap {\mathbb T}^2$ is the interior of the cylinder
$\mbox{l\hspace{ -.47em}C}_i^s$.
Similarly, ${\mathbb T}^2=\partial \cT_2$ is the union of the cylinders
$\mbox{l\hspace{ -.47em}C}_1^u$ and $\mbox{l\hspace{ -.47em}C}_2^u$ bounded by $\gamma_1^u$ and $\gamma_2^u$ and
whose interiors are the intersections ${\mathbb T}^2\cap W^u(r_1)$ and
${\mathbb T}^2\cap W^u(r_2)$, respectively. See Figure~\ref{f.rectangle}.
\item \label{i.cilinders}
As a consequence of item (\ref{i.points}) in the definition of the vector field $X$, the intersection
$\overline{\mbox{l\hspace{ -.47em}C}_1^u} \cap \overline{\mbox{l\hspace{ -.47em}C}_1^s}$ is a ``rectangle"
$R$ such that
$$
R= \overline{W^s(s_1)\cap W^u(r_1)\cap
{\mathbb T}^2}
$$
and its boundary $\partial R$ of $R$ is the union of four curves
$a_1^s,\, a_2^s,\,b_1^u, \, b_2^u$ with disjoint interiors such
that
\begin{equation}\label{e.boundaryR}
a_1^s\subset \gamma_1^s, \quad a_2^s\subset \gamma_2,\quad
b_1^u\subset \gamma^u_1,\quad
b_2^u\subset \gamma^u_2.
\end{equation}
See Figure~\ref{f.rectangle}.
Note that the interiors of $b_1^u$ and $b_2^u$ are contained in
$W^s(s_1)$ and the interiors of $a_1^s$ and $a_2^s$ are contained
in $W^u(r_1)$.
\end{enumerate}
}
\end{rema}
\begin{figure}[htb]
\psfrag{s1}{$s_1$}
\psfrag{s2}{$s_2$}
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{ss}{$\sigma^s$}
\psfrag{ss+}{$\sigma^s_+$}
\psfrag{ss-}{$\sigma^s_-$}
\psfrag{su+}{$\sigma^u_+$}
\psfrag{su-}{$\sigma^u_-$}
\psfrag{g1s}{$\gamma^s_1$}
\psfrag{g2s}{$\gamma^s_2$}
\psfrag{g1u}{$\gamma^u_1$}
\psfrag{gu2}{$\gamma^u_2$}
\psfrag{b1s}{$b^s_1$}
\psfrag{b2s}{$b^s_2$}
\psfrag{C1s}{$C_1^s$}
\psfrag{C2s}{$C_2^s$}
\psfrag{b1u}{$b^u_1$}
\psfrag{b2u}{$b^u_2$}
\psfrag{R}{$R$}
\psfrag{T1}{$\cT_1$}
\psfrag{a1s}{$a^s_1$}
\psfrag{a2s}{$a^s_2$}
\includegraphics[width=11cm]{precilindrorev.eps}\hspace{1cm}
\caption{The rectangle $R$.}
\label{f.rectangle}
\end{figure}
\subsection{Partially hyperbolicity of $X$}\label{ss.partially}
We also assume that the vector field $X$ satisfies the following partially hyperbolic
conditions:
\begin{enumerate}
\item[(1)] There is a partially hyperbolic splitting of $X$ over the
circle $\sigma^s$ (recall \eqref{i.circles} in Section~\ref{ss.auxiliary})
of the form
$$
T_{\sigma^s}\SS^3= E^{ss}\oplus E^{cs},
$$
where $E^{cs}$ is a $2$-dimensional central bundle containing the
$X$ direction, and $E^{ss}$ is a strong stable bundle that is
oriented along the circle $\sigma^s$.
A similar condition holds for the circle $\sigma^u$: There is a
partially hyperbolic splitting of $X$ over $\sigma^u$
of the form
$$
T_{\sigma^u}\SS^3= E^{cu} \oplus E^{uu},
$$
where $E^{cu}$ is a $2$-dimensional center bundle
containing the $X$-direction
and $E^{uu}$ is
a strong unstable bundle that is oriented along $\sigma^u$.
\item[(2)]
Consider the two-dimensional strong stable manifold $W^{ss}(\sigma^s)$ of $\sigma^s$ that is tangent to
$X\oplus E^{ss}$ along $\sigma^s$. Define the
local strong stable
manifold of $\sigma^s$ by $W^{ss}_{\operatorname{loc}} (\sigma^s)=W^{ss} (\sigma^s) \cap \cT_1$.
Then the intersection between $W^{ss}_{\operatorname{loc}} (\sigma^s)$
and ${\mathbb T}^2$ consists of two disjoint
$\cT_1$-parallels $\sigma^s_-$ and $\sigma^s_+$. Similarly, the intersection
$W^{uu}_{\operatorname{loc}} (\sigma^u)\cap
{\mathbb T}^2$ is the disjoint union of two $\cT_2$-parallels $\sigma^u_+$ and
$\sigma^u_-$.
We require that
$$
\mbox{l\hspace{ -.47em}C}_1^u \cap \big( \sigma^s_- \cup \sigma^s_+ \big)=\emptyset \quad
\mbox{and} \quad \mbox{l\hspace{ -.47em}C}_1^s \cap \big( \sigma^u_- \cup \sigma^u_+
\big)=\emptyset.
$$
\end{enumerate}
\begin{figure}[htb]
\psfrag{s1}{$s_1$}
\psfrag{s2}{$s_2$}
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{q1}{$q_1$}
\psfrag{q2}{$q_2$}
\psfrag{ss}{$\sigma^s$}
\psfrag{ss+}{$\sigma^s_+$}
\psfrag{ss-}{$\sigma^s_-$}
\psfrag{su+}{$\sigma^u_+$}
\psfrag{su-}{$\sigma^u_-$}
\psfrag{g1s}{$\gamma^s_1$}
\psfrag{g2s}{$\gamma^s_2$}
\psfrag{g1u}{$\gamma^u_1$}
\psfrag{gu2}{$\gamma^u_2$}
\psfrag{b1s}{$b^s_1$}
\psfrag{b2s}{$b^s_2$}
\psfrag{b1u}{$b^u_1$}
\psfrag{b2u}{$b^u_2$}
\psfrag{R}{$R$}
\psfrag{T1}{$\cT_1$}
\psfrag{a1s}{$a^s_1$}
\psfrag{a2s}{$a^s_2$}
\includegraphics[width=12cm]{strongrevb.eps}\hspace{1cm}
\caption{The rectangle $R$ and the ``strong'' manifolds.}
\label{f.strong}
\end{figure}
Let us explain how this property can be obtained. Recall that
$W^u(r_1)\cap {\mathbb T}^2$ is the interior of the cylinder $\mbox{l\hspace{ -.47em}C}_1^u$
bounded by $\gamma_1^u$ and
$\gamma_2^u$. Thus, since $\gamma_1^u$, $\gamma_2^u$,
$\sigma_-^s,$ and $\sigma_+^s$ are $\cT_1$-parallels (or
equivalently $\cT_2$-meridians), we can assume that $\mbox{l\hspace{ -.47em}C}_1^u \cap
\big( \sigma^s_- \cup \sigma^s_+ \big)=\emptyset$. See
Figure~\ref{f.strong}. The condition for the cylinder $\mbox{l\hspace{ -.47em}C}_1^s$
and the circles $\sigma^u_+$ and $\sigma^u_-$ follows identically
noting that $\gamma_1^s$, $\gamma_2^s$, $\sigma_-^u$, and
$\sigma_+^u$ are $\cT_2$-parallels.
By the partially hyperbolic conditions, the strong stable
manifolds $W^{ss}(p_i)$, $i=1,2$, (tangent to $E^{ss}$ at $p_i$) are well defined and
has dimension one. Similarly, the strong unstable manifold
$W^{uu}(q_1)$ and $W^{uu}(q_2)$ are well defined and have dimension
one. As a consequence of item (2) above (see Figure~\ref{f.strong}) we have
the following:
\begin{equation}\label{e.boundaryRbis}
\begin{split}
& a_1^s\cap W^{ss}(p_1)=\emptyset, \quad a_2^s\cap
W^{ss}(p_2)=\emptyset,\\ & b_1^u\cap W^{uu}(q_1) =\emptyset, \quad
b_2^u\cap W^u(q_2) =\emptyset.
\end{split}
\end{equation}
\subsection{Transverse heteroclinic intersection}
Consider the ``corner'' points of the rectangle $R$,
\begin{equation}\label{e.xij}
a_1^s\cap b_1^u\stackrel{\scriptscriptstyle\rm def}{=} x_{1,1}, \quad a_1^s \cap b_2^u \stackrel{\scriptscriptstyle\rm def}{=}
x_{1,2}, \quad a_2^s\cap b_1^u\stackrel{\scriptscriptstyle\rm def}{=} x_{2,1}, \quad a_2^s \cap
b_2^u\stackrel{\scriptscriptstyle\rm def}{=} x_{2,2}.
\end{equation}
By definition, the $\omega$ and $\alpha$-limits of the point $x_{i,j}$ are
the singularities $p_i$ and $q_j$, respectively.
Denote by $\Gamma_{i,j}$ the closure of the orbit of $x_{i,j}$ (see
Figures~\ref{f.gaij} and \ref{f.gaijbis}). As the intersection between $W^s(p_i)$ and
$W^u(q_j)$ is exactly the orbit of $x_{i,j}$, we have
\begin{equation}\label{e.curvasGa}
\Gamma_{i,j}\stackrel{\scriptscriptstyle\rm def}{=} \overline{W^s(p_i)\cap W^u(q_j)}=\{p_i\} \cup
\{q_j\} \cup \big(W^s(p_i) \pitchfork W^u(q_j) \big) .
\end{equation}
\begin{figure}[htb]
\psfrag{s1}{$s_1$}
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{q1}{$q_1$}
\psfrag{q2}{$q_2$}
\psfrag{G11}{$\Gamma_{1,1}$}
\psfrag{G12}{$\Gamma_{1,2}$}
\psfrag{G21}{$\Gamma_{2,1}$}
\psfrag{G22}{$\Gamma_{2,2}$}
\includegraphics[width=6cm]{curvesGadijon}\hspace{1cm}
\caption{The curves $\Gamma_{i,j}$. Global Dynamics}
\label{f.gaij}
\end{figure}
\begin{figure}[htb]
\psfrag{s1}{$s_1$}
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{q1}{$q_1$}
\psfrag{q2}{$q_2$}
\psfrag{G11}{$\Gamma_{1,1}$}
\psfrag{G12}{$\Gamma_{1,2}$}
\psfrag{G21}{$\Gamma_{2,1}$}
\psfrag{G22}{$\Gamma_{2,2}$}
\includegraphics[width=6cm]{outline}\hspace{1cm}
\caption{Outline of the curves $\Gamma_{i,j}$.}
\label{f.gaijbis}
\end{figure}
\begin{rema}\label{r.Gaij}
{\em{The curve $\Gamma_{i,j}$ is a $C^1$-invariant normally
hyperbolic compact segment.}}
\end{rema}
This remark is a standard consequence of the following facts:
\begin{itemize}
\item
The point $x_{i,j}$ is a transverse heteroclinic intersection associated
to the singularities $p_i$ and $q_j$.
\item
The partial hyperbolicity hypothesis at the singularities
$p_i$ and $q_j$
implies that $W^{ss}(p_i)$ and
$W^{uu}(q_j)$ are well defined.
\item
By construction, recall equation~\eqref{e.boundaryRbis},
$x_{i,j}\not\in W^{ss}(p_i) \cup
W^{uu}(q_j)$.
\item
The curve $\Gamma_{i,j}$ is the closure of the orbit of $x_{i,j}$.
\end{itemize}
\subsection{Invariant manifolds of the segments $\Gamma_{i,j}$}\label{ss.invariant}
For each singularity $q_j$, we have that $\big( W^u(q_j)\setminus
W^{uu}(q_j) \big)$ is the disjoint union of two connected
invariant surfaces
$$
\big( W^u(q_j)\setminus W^{uu}(q_j) \big) \stackrel{\scriptscriptstyle\rm def}{=} W^{u,+}(q_j)\cup
W^{u,-} (q_j),
$$
where $W^{u,+}(q_j)$ contains the interior of the
curves $\Gamma_{1,j}$ and $\Gamma_{2,j}$. With this notation, the
invariant manifolds of the curve $\Gamma_{i,j}$ are
\[
\begin{split}
&W^{u}(\Gamma_{i,j})= W^u(p_i) \cup W^{u,+}(q_j) \cup
W^{uu}(q_j),
\\
&W^{s}(\Gamma_{i,j})= W^s(q_j) \cup W^{s,+}(p_i) \cup
W^{ss}(p_i).
\end{split}
\]
See Figures~\ref{f.gaijbis} and \ref{f.unstable}. Note that these manifolds are
injective $C^1$-immersions of $[0,1]\times \mbox{l\hspace{ -.15em}R}$.
\begin{figure}[htb]
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{q1}{$q_1$}
\psfrag{G11}{$\Gamma_{1,1}$}
\psfrag{g21}{$\Gamma_{2,1}$}
\psfrag{Wup1}{$W^u(p_1)$}
\psfrag{Wup2}{$W^u(p_2)$}
\psfrag{wuuq1}{$W^{uu}(q_1)$}
\psfrag{wu+q1}{$W^{u,+}(q_1)$}
\psfrag{wu-q1}{$W^{u,-}(q_1)$}
\psfrag{s1}{$s_1$}
\psfrag{b1u}{$b^{u}_1$}
\psfrag{x11}{$x_{1,1}$}
\psfrag{x21}{$x_{2,1}$}
\includegraphics[width=7cm]{unstablerev.eps}\hspace{1cm}
\caption{The unstable manifold of $\Gamma_{i,j}$.} \label{f.unstable}
\end{figure}
Figures~\ref{f.gaij}, \ref{f.gaijbis}, and \ref{f.unstable} suggest that the curves $\Gamma_{1,j}$
and $\Gamma_{2,j}$ form a ``cusp" at the point $q_j$. This geometric configuration
will play a key role in our construction. So let us define precisely
what we mean by a {\emph{cusp.}}
A topological two-disk $\GG$ contained in the interior of a smooth surface $S$
{\emph{has a cusp}} at a point $p\in\partial \GG$ if for every
$\varepsilon
>0$ there are a convex cone $C_\varepsilon$ of angle $\varepsilon$
at $p$ and a neighborhood $U_\varepsilon$ of $p$ in $S$ such that
$\GG\cap U_\varepsilon\subset C_\varepsilon$. Given a curve
$\gamma$ contained in the interior of the surface $S$, a point $q$ in the interior of
$\gamma$ is a {\emph{cusp}} of
$\gamma$ if there is a topological disk $\mbox{I\hspace{ -.15em}F}$ whose boundary
contains $\gamma$ and has a cusp at $q$.
\begin{rema} \label{r.curvesGaij}{\em{
Recall that the interior of the curves $\Gamma_{1,i}$ and
$\Gamma_{2,i}$ are disjoint from $W^{uu}(q_i)$. Moreover, the
interior of these curves are the orbits of the points $x_{1,i}$
and $x_{2,i}$, respectively. These two curves are connected by the
segment $b_i^u\subset W^u(q_i)$ that is disjoint from $W^{uu}(q_i)$,
recall \eqref{e.boundaryR} and see Figure~\ref{f.unstable}. The partial hyperbolicity hypothesis
now implies that $\Gamma_{1,i}$ and $\Gamma_{2,i}$ are ``central curves''
arriving to $q_1$ from the same side of $W^{uu}(q_i)$. These two
conditions imply that
\begin{equation}
\label{e.Gaqi}
\Gamma (q_i)\stackrel{\scriptscriptstyle\rm def}{=} \Gamma_{1,i}\cap \Gamma_{2,i}, \quad i=1,2,
\end{equation}
is a curve with a \emph{cusp singularity} at $q_1$, see Figures~\ref{f.gaij},
\ref{f.gaijbis},
and
\ref{f.unstable}.}}
\end{rema}
The unstable manifold $W^u(\Gamma(q_i))$ of $\Gamma(q_i)$ is the set
$W^u(\Gamma_{1,i})\cup W^u(\Gamma_{2,i})$. Noting that the interior of
the ``strip" $W^u(\Gamma(q_i))$ is $W^{u,+}(q_i)$, $i=1,2$, we get
\begin{equation}
\label{e.q} W^u(\Gamma (q_i))= W^u(p_1)\cup W^u(p_2) \cup
W^{u,+}(q_i) \cup W^{uu}(q_i).
\end{equation}
We observe that the set $W^u(\Gamma (q_i))$ is an injective
$C^1$-immersion of a connected surface with boundary. Equivalent
statements hold for
\begin{equation}
\label{e.Gapi}
\Gamma(p_i)\stackrel{\scriptscriptstyle\rm def}{=} \Gamma_{i,1}\cup \Gamma_{i,2}, \quad i=1,2,
\end{equation}
and its stable manifold
\begin{equation}
\label{e.p} W^s(\Gamma(p_i))= W^s(q_1)\cup W^s(q_2) \cup W^{s,+}(p_i)
\cup W^{ss}(p_i),
\end{equation}
where $W^{s,+}(p_i)$ is the component of
$\big( W^s(p_i)\setminus
W^{ss}(p_i) \big)$ containing the interior of the curves
$\Gamma_{i,1}$ and $\Gamma_{i,2}$.
With this notation, the sides $a_i^s$ and $b_j^u$
of the rectangle $R$ (see \eqref{e.boundaryR}) satisfy the
following property
\begin{equation} \label{e.sides}
a_i^s \subset W^s(\Gamma(p_i))\cap {\mathbb T}^2 \quad
\mbox{and}\quad b_j^u\subset W^u(\Gamma(q_j))\cap {\mathbb T}^2, \qquad
i,j\in\{1,2\}.
\end{equation}
\subsection{Central bundles}
\label{ss.central}
In this section, we see that the unstable manifolds of $\Gamma_{i,1}$ and
$\Gamma_{i,2}$ touch each other
at $W^u(p_i)$ tangentially ``coming from the same side" of
$W^u(p_i)$, see Figure~\ref{f.ecplus}. In the following we will precise
what this means.
\begin{lemm}[Center stable/unstable
bundles]\label{l.centerbundles}
Given any singularity $p$ of
saddle type with a strong stable direction $W^{ss}(p)$ (tangent
to a strong stable bundle $E^{ss}(p)$) there is a unique invariant
``central" bundle $E^c$ defined over the unstable manifold $W^u(p)$ of $p$ that is transverse at
$p$ to the bundle $E^{ss}$ and has codimension $\dim (E^{ss})$.
A similar property holds for saddle singularities $q$ with a
strong unstable manifold tangent to some strong unstable bundle
$E^{uu}$. In this case, there is a central bundle
defined over
$W^s(q)$ that is
transverse to $E^{uu}$ at $q$ and has codimension $\dim (E^{uu})$.
\end{lemm}
\begin{proof}
We can assume that for every point $x\in W^u_{{\operatorname{loc}}}(p)$
there
is defined a
negatively invariant cone-field $\cC^{ss}$ around the strong
stable direction $E^{ss}$. Consider the complement $\cC^c$ of
$\cC^{ss}$. Given $y\in W^u(p)$ there is $t(y)>0$ such that
$X_{-t}(y) \in W^u_{{\operatorname{loc}}}(p)$ for all $t\ge t(y)$.
Given $y\in W^u(p)$ it is enough to define
$$
E^{c}(y)\stackrel{\scriptscriptstyle\rm def}{=} \{ v\, \colon \, D_y X_{-t}(v) \in
\cC^{c}(X_{-t}(y)) \quad \mbox{for all $t\ge t(y)$}\}.
$$
By construction, the bundle $E^c(y)$ is transverse to $E^{ss}$ and its dimension
is the codimension of $E^{ss}$.
This completes the proof of the lemma.
\end{proof}
\begin{rema}\label{r.notation}
{\em{
Applying Lemma~\ref{l.centerbundles} to the singularity $p_i$, we get a two
dimensional bundle $E^c(p_i)$, $i=1,2$, coinciding with the
bundle $E^{cs}$ defined along the curve $\sigma^s$ in
Section~\ref{ss.partially}. Analogously, the bundle $E^c$ defined
along $W^s(q_j)$ coincides with the bundle $E^{cu}$ along
$\sigma^u$.}}
\end{rema}
\begin{lemm}\label{l.tangent}
The surfaces $W^u(\Gamma(q_1))$ and $W^u(\Gamma(q_2))$ are tangent to the bundle
$E^{cs}$ along
their intersection $W^u(p_1)\cup W^u(p_2)$. Similarly, the
surfaces $W^s(\Gamma(p_1))$ and $W^s(\Gamma(p_2))$ are tangent
to the bundle $E^{cu}$
along their intersection $W^s(q_1)\cup
W^s(q_2)$.
\end{lemm}
\begin{proof}
The boundary part of $W^u(\Gamma(q_j))$ has three components,
$W^u(p_1)$, $W^u(p_2)$, and $W^{uu}(q_j)$,
recall equation \eqref{e.q} and see Figure~\ref{f.unstable}. Moreover, the
surface $W^u(\Gamma(q_j))$ is transverse to $W^{ss}(p_i)$. The
uniqueness of the central bundle $E^c$ in
Lemma~\ref{l.centerbundles} implies that for each
$x\in W^u(p_i)\subset \sigma^s$
the fiber $E^c(x)=E^{cs}(x)$ (recall Remark~\ref{r.notation}) is the tangent space $T_x(W^u(\Gamma(q_j)))$. This
implies the lemma.
\end{proof}
\begin{figure}[htb]
\psfrag{p1}{$p_i$}
\psfrag{q2}{$q_2$}
\psfrag{q1}{$q_1$}
\psfrag{wuG11}{$W^u(\Gamma_{i,1})$}
\psfrag{wuG12}{$W^u(\Gamma_{i,2})$}
\psfrag{wssp1}{$W^{ss}(p_i)$}
\psfrag{G11}{$\Gamma_{i,1}$}
\psfrag{G12}{$\Gamma_{i,2}$}
\psfrag{Ec+}{$E^{cs}_+$}
\psfrag{Ec}{$E^{cs}$}
\includegraphics[width=7cm]{Ecplus.eps}\hspace{1cm}
\caption{The open ``half-planes" $E^{cu}_{+}$.}
\label{f.ecplus}
\end{figure}
\begin{rema}[The open ``half-planes" $E^{cs}_{+}$ and $E^{cu}_{+}$]
\label{r.semiplanes} {\em{ The normally hyperbolic curves
$\Gamma_{i,1}$ and $\Gamma_{i,2}$ are ``central curves" contained in
$W^s(p_i)$ arriving to $p_i$ from the same side of $W^s(p_i)\setminus
W^{ss}(p_i)$.
Furthermore, the boundary surfaces $W^u(\Gamma_{i,1})$ and
$W^u(\Gamma_{i,2})$ are tangent to $E^{cs}$ along $W^u(p_i)$. This
implies that, for every $x\in W^u(p_i)$, the vectors in $E^{cs}(x)$
pointing to the interior of $W^u(\Gamma(q_1))$) form an open half-plane
$E^{cs}_+(x)$. This half-plane coincides with the vectors of $E^{cs}(x)$
pointing to the interior of
$W^u(\Gamma_{i,1})$ or (equivalently)
of
$W^u(\Gamma_{i,2})$. See Figure~\ref{f.ecplus}.
For points $y\in W^s(q_j)$, we similarly define the half-plane
$E^{cu}_+(y)$ as the vectors in $E^{cu}$ pointing to the interior of
$W^s(\Gamma(p_i))$, $i=1,2$.}}
\end{rema}
\subsection{Position of the invariant manifolds in the basins of $r_1$ and
$s_1$}\label{s.positionofinvariant} Consider a ``small" two-sphere
$\SS^s$ contained in the interior of the solid torus $\cT_1$ that
is transverse to the vector field $X$ and bounds a three-ball
$\BB^s\subset W^s(s_1) \cap \cT_1$ whose interior contains the
singularity $s_1$. Let
\begin{equation}\label{e.etas}
\eta^s\stackrel{\scriptscriptstyle\rm def}{=} \SS^s \cap W^{ss}(\sigma^s).
\end{equation}
Note that $\eta^s$ is a circle that contains the points
\begin{equation}\label{e.yu12}
y_1^u\stackrel{\scriptscriptstyle\rm def}{=} W^u(p_1) \cap \SS^s \quad \mbox{and} \quad y_2^u\stackrel{\scriptscriptstyle\rm def}{=}
W^u(p_2) \cap \SS^s, \qquad y_1^u,y^u_2\in \eta^s.
\end{equation}
We similarly define a ``small" two-sphere $\SS^u \subset \cT_2$
transverse to $X$ bounding a three-ball $\BB^u\subset W^u(r_1)$ whose
interior contains $r_1$. We define the
circle
$$
\eta^u\stackrel{\scriptscriptstyle\rm def}{=}
\SS^u\cap
W^{uu}(\sigma^u)
$$
and the intersection
points
\begin{equation}\label{e.ys12}
y_1^{s}\stackrel{\scriptscriptstyle\rm def}{=} W^s(q_1) \cap \SS^u \quad \mbox{and} \quad
y^s_2= W^s(q_2) \cap \SS^u, \qquad y_1^s,y^s_2\in \eta^u.
\end{equation}
\begin{rema}\label{r.time}
{\em{Choosing the balls $\BB^s$ and $\BB^u$ small enough, we can assume that
the minimum time that a point takes to go from $\BB^u$ to $\BB^s$ is arbitrarily large. In particular, this time is bigger than $10$:
$X_t (\BB^u) \cap \BB^s=\emptyset$ for all $t\in [0,10]$. }}
\end{rema}
Consider the sets $\GG^u$ and $\GG^s$ (the set $\GG^ u$ is depicted
in Figure~\ref{f.Gu}),
$$
\GG^u \stackrel{\scriptscriptstyle\rm def}{=} \overline{W^u(r_1) \cap \SS^s} \quad \mbox{and} \quad
\GG^s\stackrel{\scriptscriptstyle\rm def}{=} \overline{W^s(s_1) \cap \SS^u}.
$$
\begin{figure}[htb]
\psfrag{p1}{$p_1$}
\psfrag{p2}{$p_2$}
\psfrag{s1}{$s_1$}
\psfrag{q1}{$q_1$}
\psfrag{q2}{$q_2$}
\psfrag{l1u}{$\ell_{1}^u$}
\psfrag{l2u}{$\ell_{2}^u$}
\psfrag{y1u}{$y_{1}^u$}
\psfrag{y2u}{$y_{2}^u$}
\psfrag{G1}{$\GG^{u}$}
\psfrag{Ss}{$\SS^{s}$}
\includegraphics[width=9cm]{cuspdijon.eps}\hspace{1cm}
\caption{The set $\GG^u$.}
\label{f.Gu}
\end{figure}
\begin{lemm}\label{l.banana}
The set $\GG^u$ is a topological two-disk bounded by
$ \ell_1^u \cup \ell_2^u \cup \{y_1^u\} \cup \{y_2^u\}$, where
$$
\ell_1^u \subset W^u(q_1) \quad \mbox{and} \quad \ell_2^u \subset
W^u(q_2)
$$
are disjoint (open) simple curves whose endpoints are $y_1^u$ and
$y_2^u$. The closed curves $\overline{\ell_1^u}$ and
$\overline{\ell_2^u}$ have the same tangent direction at their
endpoints $y^u_1$ and $y^u_2$. The disk $\GG^u$ has two cusps at
the points
$y^u_1$ and $y^u_2$.
Similarly,
the set $\GG^s$ is a topological two-disk bounded by $\ell_1^s \cup
\ell_2^s \cup \{y_1^s\} \cup \{y_2^s\}$, where
$$
\ell_1^s \subset W^s(p_1) \quad \mbox{and} \quad \ell_2^s \subset
W^s(p_2)
$$
are disjoint (open) simple curves whose endpoints are $y_1^s$ and
$y_2^s$. The closed curves $\overline{\ell_1^s}$ and
$\overline{\ell_2^s}$ have the same tangent direction at their
endpoints $y^s_1$ and $y^s_2$.
The disk $\GG^s$ has two cusps at $y_1^s$ and $y^s_2$.
\end{lemm}
We consider the following notation, given an interval $[t_1,t_2]$
and a set $A$ we let
$$
X_{[t_1,t_2]}(A)\stackrel{\scriptscriptstyle\rm def}{=} \bigcup_{t\in [t_1,t_2]} X_t(A).
$$
\begin{proof} We only prove the lemma for the set $\GG^u$, the proof
for $\GG^s$ is identical.
Note that the sphere $\SS^s$ intersects every orbit of the set
$W^s(s_1)\setminus \{s_1\}$ in exactly one point.
Thus, since
the rectangle $R$ in Remark~\ref{r.stablecylinders} is the closure
of $W^u(r_1)\cap W^s(s_1)\cap {\mathbb T}^2$, the
positive orbit of any point in the interior of $R$ intersects
$\SS^s$ in exactly one point. Hence the set $\GG^u$ is the
closure of the ``projection" along the orbits of $X$ of the
interior of $R$ into $\SS^s$, that is,
$$
\GG^u= \overline{
X_{[0,\infty)} \big( \mbox{int}(R) \big) \cap \SS^s}.
$$
By construction the set $\GG^u$ is a topological two-disk. We next
describe its boundary.
By equation \eqref{e.boundaryR} the boundary of $R$ consists of
the segments $a_i^s\subset W^s(p_i)$ and $b_i^u\subset W^u(q_j)$,
$i=1,2$. Furthermore, the interiors of the segments $b_1^u$ and
$b_2^u$ are contained in $W^s(s_1)$. Denote by $\ell^u_1$ and
$\ell^u_2$ the ``projections'' by the flow of $X$ of these interiors
into $\SS^s$, that is,
$$
\ell_i^u \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,\infty)} \big( \mbox{int} (b_i^u) \big) \cap
\SS^s.
$$
Consider any sequence $(x_n)$ of points in the interior of $R$
accumulating to the side $a_i^s$ of $R$. Note that the (positive)
orbit of $x_n$ by the flow of $X$ goes arbitrarily close to the
saddle singularity $p_i$ before intersecting
$\SS^s$ at a point $y_n$. By construction, the sequence $(y_n)$ converges
to $y_i^u=W^u(p_i)\cap \SS^s$. Indeed, for any given curve $b\subset R$ transverse to $X$ joining the
sides $a_1^s$ and $a_2^s$ of $R$ the intersection of the sphere $\SS^s$
and the positive orbit of $b$ by the flow
$X$ (i.e., the
``projection" of $b$ into $\SS^s$ by the flow)
is a
curve $\ell_b$ joining $y^u_2$ and $y_2^u$ (these points are in
the closure of $\ell_b$). In particular, $y_1^u$ and $y_1^u$ are
the endpoints of $\ell^u_i$, $i=1,2$.
Bearing in mind equation~\eqref{e.sides}
and the definitions of $y_{i}^u$, $\ell^u_{i}$, and $\Gamma(q_i)$,
$i=1,2$, we get the following:
\[
\begin{split}
\overline{\ell^u_i} &= \ell^u_i
\cup \{y_1^u,y_2^u\}= W^u(\Gamma(q_i)) \cap \SS^s,
\\
\partial \GG^u&= \big( W^u(\Gamma(q_1)) \cap \SS^s \big) \cup \big( W^u(\Gamma(q_2)) \cap \SS^s
\big)= \ell^u_1 \cup \ell^u_2 \cup \{y_1^u,y_2^u\}.
\end{split}
\]
This completes the description of the set $\partial
\GG^u$.
It remains to see that $y_1^u$ and $y_2^u$ are cusps of $\GG^u$.
By Lemma~\ref{l.tangent}
the curves $\overline{\ell^u_1}$ and
$\overline{\ell^u_2}$ are tangent at $y_i^u$ to
$E^{cs}(y_i^u)\cap T_{y_i^u}(\SS^s)$.
To see that the point $y_i^u$ is a cusp of $\GG^u$
it is enough to note that the interior of $\GG^u$ is disjoint from the circle
$\eta^s$.
Thus the disk $\GG^u$ is the ``thin component" of $\SS^s\setminus
\partial \GG^u$. This ends the proof of the lemma.
\end{proof}
\begin{rema}
\label{r.lastinclusion}
With the notations above, the following inclusions hold
$$
(\SS^s\setminus \GG^u) \subset W^u(r_2) \quad \mbox{and} \quad
(\SS^u\setminus \GG^s\subset W^s(s_2)).
$$
\end{rema}
\subsection{The diffeomorphism time-one map $X_1$}\label{ss.gluing}
Let $X_1$ denote the time-one map of the vector field $X$ and
define the diffeomorphism $F_0\stackrel{\scriptscriptstyle\rm def}{=} X_1.$ Note that $F_0$ is a
Morse-Smale diffeomorphism whose non-wandering set consists of the
sinks $s_1$ and $s_2$, the saddles of $s$-index two $p_1$ and
$p_2$, the saddles of $s$-index one $q_1$ and $q_2$, and the
sources $r_1$ and $r_2$. Note that the invariant manifolds of these
points for the vector field $X$ and for the diffeomorphism $F_0$ coincide. We only write
$W^i(x,X)$ or $W^i(x,F_0)$ to emphasize the role of $X$ or $F_0$,
otherwise we just write $W^i(x)$.
Consider the fundamental domain $\Delta^s$ of $W^s(s_1)$ for $F_0$
bounded by $\SS^s$ and $F_0(\SS^s)$. Note that
$$
\Delta^s \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (\SS^s) = \BB^s\setminus
\mbox{int}(F_0(\BB^s)) \simeq \SS^s\times [0,1].
$$
Let
$$
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u\stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (\GG^u), \quad {\mathbb L}^u_i \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]}
(\ell^u_i), \quad {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (y^u_i).
$$
These sets are depicted in
Figure~\ref{f.Eu}.
\begin{figure}[htb]
\psfrag{l1u}{$\ell_{1}^u$}
\psfrag{l2u}{$\ell_{2}^u$}
\psfrag{y1u}{$y_{1}^u$}
\psfrag{y2u}{$y_{2}^u$}
\psfrag{L1u}{$L_{1}^u$}
\psfrag{L2u}{$L_{2}^u$}
\psfrag{Y1u}{$Y_{1}^u$}
\psfrag{Y2u}{$Y_{2}^u$}
\psfrag{Gu}{$\GG^{u}$}
\psfrag{FGu}{$F_0(\GG^{u})$}
\psfrag{Eu}{${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{u}$}
\psfrag{G1}{$\GG^{u}$}
\psfrag{Ss}{$\SS^{s}$}
\psfrag{Ds}{$\Delta^{s}$}
\includegraphics[width=9cm]{Eu.eps}\hspace{1cm}
\caption{The sets ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u, {\mathbb L}_i^u,$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i$.}
\label{f.Eu}
\end{figure}
By construction, the set ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u$ is a topological ball bounded by the disks $\GG^u$ and
$F_0(\GG^u)$, the ``rectangles" ${\mathbb L}_1^u$ and ${\mathbb L}^u_2$, and
the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2$. By the definition of $\GG^u$
and of the cylinders $\mbox{l\hspace{ -.47em}C}_1^u$ and $\mbox{l\hspace{ -.47em}C}^u_2$ and by
Lemma~\ref{l.banana} (see also Remark~\ref{r.lastinclusion}), we have that
\begin{equation} \label{e.inEs}
\overline{W^u (r_1) \cap
\Delta^s}={\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u \quad \mbox{and} \quad \Delta^s\setminus {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u \subset
W^u(r_2).
\end{equation}
Since $\ell^u_i\subset W^u(q_i,X)$ and $y^u_i\subset W^u(p_i,X)$
we have
\begin{equation}\label{e.LsYs}
{\mathbb L}^u_i \subset W^u(q_i,F_0) \quad \mbox{and} \quad
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i \subset W^u(p_i,F_0).
\end{equation}
Note that the common boundary of the rectangles ${\mathbb L}^u_1$ and
${\mathbb L}^u_2$ are the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_1^u$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2$. Moreover,
the closure of ${\mathbb L}_i^u$ is tangent to the center unstable bundle
$E^{cu}$ along the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_1^u$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2$.
We similarly define the fundamental domain $\Delta^u$ of $W^u(r_1)$
for $F_0$ bounded by $\SS^u$ and $F_0(\SS^u)$. As before
$$
\Delta^u \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (\SS^u) \simeq \SS^u\times [0,1].
$$
We let
$$
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s\stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (\GG^s), \quad {\mathbb L}^s_i \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]}
(\ell^s_i), \quad {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_i \stackrel{\scriptscriptstyle\rm def}{=} X_{[0,1]} (y^s_i)
$$
and as above we have
\begin{equation}\label{e.LuYu}
{\mathbb L}^s_i \subset W^s(p_i) \quad \mbox{and} \quad
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_i \subset W^s(q_1).
\end{equation}
Moreover, the ``rectangles" ${\mathbb L}^s_1$ and ${\mathbb L}^s_2$ are tangent
at the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_1^s$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_2$ to the center stable bundle
$E^{cs}$. By construction, ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s$ is bounded by the disks $\GG^s$
and $F_0(\GG^s)$, the ``rectangles" ${\mathbb L}_1^s$ and ${\mathbb L}^s_s$, and
the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_2$. Finally,
\begin{equation}
\label{e.inEu} \overline{W^s (s_1) \cap \Delta^u}={\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s \quad
\mbox{and} \quad \Delta^u\setminus {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s \subset
W^s(s_2).
\end{equation}
\section{A diffeomorphism on $\SS^2\times \SS^1$ obtained by a
surgery}\label{s.surgery}
\subsection{The surgery}\label{ss.surgery}
In this section
identify some regions of $\SS^3$ by a local
diffeomorphism $\Psi$. This surgery provides
a diffeomorphism $F_\Psi$ defined on $\SS^2 \times \SS^1$
induced by $F_0=X_1$ and the quotient of $\SS^3$ by $\Psi$.
We will see in Section~\ref{s.psi}
that for an appropriate choice of $\Psi$ the diffeomorphism $F_\Psi$
has fragile cycles.
With the notation in Section~\ref{s.auxiliary},
consider a diffeomorphism $\Psi\colon \Delta^s\to \Delta^u$
such that
\begin{itemize}
\item
$\Psi(\SS^s)= \SS^u$ and $\Psi(F_0(\SS^s))=F_0(\SS^u)$,
\item
$\Psi\circ F_0|_{\SS^s}= F_0|_{\SS^u}\circ \Psi$,
\item
$D\Psi\circ DF_0|_{\SS^s}= DF_0|_{\SS^u}\circ D\Psi$.
\end{itemize}
Recall that $\BB^s\subset W^s(s_1)$ and $\BB^u\subset W^u(r_1)$
are small balls containing $s_1$ and $r_1$, respectively.
In the set $ \SS^3\setminus \Big( \mbox{int}
\big( F_0(\BB^s) \big)\cup \mbox{int}\big( \BB^u \big) \Big)$ we
identify the
points $x\in \Delta^s$ and $\Psi(x)\in \Delta^u$ obtaining the
quotient space
$$
M\stackrel{\scriptscriptstyle\rm def}{=} \left( \SS^3\setminus \Big( \mbox{int} \big( F_0(\BB^s)
\big)\cup \mbox{int}\big( \BB^u) \big)\Big )\right) \Big\slash
\Psi.
$$
The set $M$ is a $C^1$-manifold diffeomorphic to $\SS^2\times \SS^1$
and the diffeomorphism $F_0$ induces a diffeomorphism $F_\Psi
\colon M\to M$.
Denote by $\pi$ the projection $\pi\colon \SS^3\setminus \Big(
\mbox{int} \big( F_0(\BB^s) \big)\cup \mbox{int}\big( \BB^u)
\big)\Big ) \to M$ that associates to $x$ its class
$\pi(x)=[x]$ by the equivalence relation induced by $\Psi$. For
notational simplicity, if $x\not\in
\Delta^s \cup \Delta^u$ we simply write $x$ instead of $\pi(x)$.
Write
$$
\Delta\stackrel{\scriptscriptstyle\rm def}{=} \big(\Delta^s\cup \Delta^u \big)/\Psi=\pi(\Delta^s)=\pi(\Delta^u).
$$
In what follows we write $\KK^u_i(F_\Psi)\stackrel{\scriptscriptstyle\rm def}{=} \pi (\KK^u_i)$,
where $\KK={\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z},{\mathbb L},\GG,{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}$ and $i=,1,2, \emptyset$.
\begin{rema}\label{r.delta} {\em{The set
$$
\Delta\setminus \big ({\mathbb L}_1^u (F_\Psi) \cup {\mathbb L}_2^u (F_\Psi)\cup
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F_\Psi) \cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2 (F_\Psi) \cup \GG^u (F_\Psi) \cup
F_\Psi ( \GG^u (F_\Psi) ) \big)
$$
has two connected components. The set ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F_\Psi)$ is the
component that is diffeomomorphic to $\GG^u \times [0,1]$. This set is a topological
ball.
There is a similar characterization for
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F_\Psi)$. }}
\end{rema}
\subsection{The dynamics of $F_\Psi$}
Note that the orbits of $F_\Psi$ disjoint from $\Delta$ are the
projection of the orbits of the diffeomorphism $F_0$.
Thus $s_2$ is a sink and $r_2$ is a source of $F_\Psi$
(note that the sink $s_1$ and the source $r_1$ are removed
in the construction of $M$). Similarly, the points $p_1$ and $p_2$ are saddles
of $s$-index $2$ of $F_\Psi$ and the points $q_1$ and $q_2$ are saddles of $s$-index
$1$ of $F_\Psi$. Observe that $F_{\Psi}$
can have further periodic points, but by
Remark~\ref{r.time}
these points
have period larger
than $10$.
Using the identification by $\Psi$ and the properties of $F_0$,
we
have the following characterization of the sets ${\mathbb L}^{u,s}_i(F_\Psi)$
and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^{u,s}_i(F_\Psi)$, $i=1,2$:
\begin{equation}\label{e.inclusions}
\begin{split}
{\mathbb L}^u_i(F_\Psi)&=\{x\in \Delta \, \colon \, x \in W^u(q_i,F_\Psi)
\,\, \mbox{and}\,\, F_\Psi^{-j}(x) \notin \Delta \,\, \mbox{for
all}\,\, j\ge
2\},\\
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i(F_\Psi)&=\{x\in \Delta \, \colon \, x \in W^u(p_i,F_\Psi)
\,\, \mbox{and}\,\, F_\Psi^{-j}(x) \notin \Delta \,\, \mbox{for
all}\,\, j\ge 2\},
\\
{\mathbb L}^s_i(F_\Psi)&=\{x\in \Delta \, \colon \, x \in W^s(p_i,F_\Psi)
\,\, \mbox{and}\,\, F_\Psi^{j}(x) \notin \Delta \,\, \mbox{for
all}\,\, j\ge 2\},
\\
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_i(F_\Psi)&=\{x\in \Delta \, \colon \, x \in W^s(q_i,F_\Psi)
\,\, \mbox{and}\,\, F_\Psi^{j}(x) \notin \Delta \,\, \mbox{for
all}\,\, j\ge 2\}.
\end{split}
\end{equation}
\begin{rema}\label{r.paredesFPsi}
{\em{
The normally hyperbolic curves $\Gamma_{i,j}$ of $F_0$ (recall Remark~\ref{r.Gaij})
are disjoint from $\BB^s\cup \BB^u$, thus their projections on $M$ (also denoted by $\Gamma_{i,j}$)
are normally hyperbolic curves of $F_\Psi$ .
Observe also that by construction,
the interior of $\Gamma_{i,j}$ is contained in
$W^s( p_i,F_\Psi) \pitchfork W^u(q_j,F_\Psi)$, recall equation
\eqref{e.curvasGa}.
We continue to use the notation
$\Gamma(q_j)=\Gamma_{1,j}\cup\Gamma_{2,j}$ and $\Gamma(p_i)=\Gamma_{i,1}\cup \Gamma_{i,2}$.
}}
\end{rema}
With the previous notation we have that
$$
{\mathbb L}^u_j(F_\Psi) \cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F_\Psi) \cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2 (F_\Psi)= \overline{{\mathbb L}^u_j(F_\Psi)}
$$
is a connected component of $W^u(\Gamma (q_j), F_\Psi)\cap \Delta$.
\begin{lemm}[Invariant manifolds and their intersections]
\label{l.bananasfirst} Consider $x\in \Delta$.
\begin{enumerate}
\item \label{banana1}
If $x\notin {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F_\Psi)$ then $x\in W^u(r_2,F_\Psi)$ and thus
it is not chain recurrent,
\item \label{banana2}
if $x\notin {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s (F_\Psi)$ then $x\in W^s(s_2,F_\Psi)$ and thus
it is not chain recurrent.
\item \label{banana3}
if $x\in {\mathbb L}^u_i (F_\Psi)$ then $x\in W^u(q_i,F_\Psi)$,
\item
\label{banana4} if
$x\in {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i (F_\Psi)$ then $x\in W^u(p_i,F_\Psi)$,
\item
\label{banana5} if $x\in {\mathbb L}^s_i (F_\Psi)$ then $x\in
W^s(p_i,F_\Psi)$, and
\item
\label{banana6} if
$x\in {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_i (F_\Psi)$ then $x\in
W^s(q_i,F_\Psi)$,
\end{enumerate}
\end{lemm}
\begin{proof}The first item follows immediately from equation \eqref{e.inEs} and the definition of $F_\Psi$.
Similarly, the second item follows from \eqref{e.inEu}.
Items (\ref{banana3})-(\ref{banana6}) follow from equation \eqref{e.inclusions}.
\end{proof}
\begin{figure}[htb]
\psfrag{z}{$z$}
\psfrag{w}{$w$}
\psfrag{Y1s}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^s(F_\Psi)$}
\psfrag{L1s}{${\mathbb L}_{1}^s(F_\Psi)$}
\psfrag{Y1u}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^u (F_\Psi)$}
\psfrag{L1u}{${\mathbb L}_{1}^u(F_\Psi)$}
\psfrag{Eu}{${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{u}(F_\Psi)$}
\psfrag{Es}{${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{s}(F_\Psi)$}
\psfrag{Ss}{$\SS^{s}$}
\psfrag{homoclinic}{$\Delta^u\cap R(F_\Psi)$}
\includegraphics[width=9cm]{lemmarev.eps}\hspace{1cm}
\caption{The recurrence region of $F_\Psi$.}
\label{f.lemma}
\end{figure}
An immediate consequence of the first two items of Lemma~\ref{l.bananasfirst}
is the following:
\begin{coro}\label{c.inextremis}
Let $x\in \Delta$ be a chain recurrent point for $F_\Psi$. Then $x\in
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s (F_\Psi)\cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F_\Psi)$.
\end{coro}
A straightforward consequence of Lemma~\ref{l.bananasfirst}
and Remark~\ref{r.paredesFPsi}
is the following:
\begin{coro}\label{c.bananas} Consider $x\in \Delta$.
\begin{enumerate}
\item \label{cbanana1}
If $x\in {\mathbb L}_i^u (F_\Psi)\cap {\mathbb L}_j^s (F_\Psi)$ then $x\in
W^u(q_i,F_\Psi) \pitchfork W^s(p_j,F_\Psi)$.
\item \label{cbanana2}
If $x\in {\mathbb L}_i^u (F_\Psi) \cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_j^s (F_\Psi)$ then $x\in
W^u(q_i,F_\Psi) \cap W^s(q_j,F_\Psi)$.
\item \label{cbanana3}
If $x\in {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_i^u (F_\Psi) \cap {\mathbb L}_j^s (F_\Psi)$ then $x\in
W^u(p_i,F_\Psi) \cap W^s(p_j,F_\Psi)$.
\item \label{cbanana4}
If $x\in {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_i^u (F_\Psi) \cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_j^s (F_\Psi)$ then $x\in
W^u(p_i,F_\Psi) \cap W^s(q_j,F_\Psi)$.
\item
\label{cbanana5} The interior of $\Gamma_{i,j}$ is contained in
$W^s(p_i,F_\Psi)\pitchfork W^u(q_j,F_\Psi)$.
\end{enumerate}
\end{coro}
Recall that a periodic point is called {\emph{isolated}} if its chain recurrent class coincides
with its (finite) orbit.
\begin{lemm}[Heterodimensional cycles and trivial homoclinic classes]\label{l.hetero}
$\,$
\begin{enumerate}
\item \label{hh1}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F_\Psi)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1 (F_\Psi)\ne\emptyset$ then
$F_\Psi$ has a heterodimensional cycle associated to $p_1$ and
$q_1$.
\item \label{hh2}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F_\Psi)\pitchfork {\mathbb L}^s_1 (F_\Psi)\neq \emptyset$
then the homoclinic class of $q_1$ is non trivial.
\item \label{hh3}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1 (F_\Psi)\pitchfork {\mathbb L}^u_1 (F_\Psi)\neq \emptyset$
then homoclinic class of $p_1$ is non trivial.
\item
\label{hh4} If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F_\Psi)\cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s (F_\Psi) =
\emptyset$ then $p_1$ is an isolated saddle.
\item
\label{hh5}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1 (F_\Psi) \cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F_\Psi) = \emptyset$ then
$q_1$ is an isolated saddle.
\end{enumerate}
\end{lemm}
\begin{proof}
To prove the first item just recall that by item (\ref{cbanana5})
in Corollary~\ref{c.bananas} the interior of $\Gamma_{1,1}$ is
contained in $W^s(p_1,F_\Psi)\pitchfork W^u(q_1,F_\Psi)$. By item
(\ref{cbanana4}) in Corollary~\ref{c.bananas}, if
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1 (F_\Psi)\ne \emptyset$ then $W^u(p_1,F_\Psi)\cap
W^s(q_1,F_\Psi)\ne\emptyset$ and thus there is a heterodimensional cycle
associated to $p_1$ and $q_1$.
Items \eqref{hh2} and \eqref{hh3} follow from Lemma~\ref{l.bananasfirst}.
To prove item \eqref{hh4}
we use the following simple fact whose proof we omit.
\begin{rema}\label{r.nonisolated}{\em{
Let $p$ be a hyperbolic saddle that is non-isolated.
Then its stable/unstable manifold contains points of its chain recurrence class
that do not belong to its orbit. In particular, if a hyperbolic fixed point $p$ is such that
$W^u(p)\setminus \{p\}$
(respectively, $W^s(p)\setminus \{p\}$)
is contained in the stable (resp. unstable) manifolds of some sinks
(resp. sources) then it is isolated.}}
\end{rema}
Note that every point $x\in W^u
(p_1,F_\Psi)$, $x\ne p_1$, in the separatrix of $W^u(p_1,F_\Psi)$
that does not contain $y_1^u$ is contained in $W^s(s_2,F_\Psi)$.
Thus this separatrix does not contain chain recurrent points.
Hence it is enough to consider points in the separatrix of
$W^u(p_1,F_{\Psi})$ containing $y_1^u$. Note that
$W^u(p_1,F_{\Psi})\cap \Delta$ contains a fundamental domain of
$W^u(p_1,F_{\Psi})$ that is contained
in ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)$.
Thus, by Remark~\ref{r.nonisolated}, if $p_1$ is not isolated then
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)$ must contain some point of the chain recurrence class of $p_1$.
Thus in such a case ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)$ cannot be contained in $W^s(s_2,F_\Psi)$.
Suppose that ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)\cap
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s (F_\Psi)= \emptyset$. Then, by item \eqref{banana2} in
Lemma~\ref{l.bananasfirst}, one has that
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)\subset W^s(s_2)$. By the above discussion
this implies that $p_1$ is isolated,
proving item
\eqref{hh4}.
The proof of item \eqref{hh5} is identical to the previous one and thus it is omitted.
\end{proof}
\section{Dynamics in a neighborhood of $F_\Psi$}
\label{ss.neighborhood}
In this section we consider
diffeomorphisms $F$ in a small $C^1$-neighborhood of $F_\Psi$.
The hyperbolic-like properties of the objects
introduced in Section~\ref{s.surgery} for the diffeomorphism $F_\Psi$
allows us to define their continuations for nearby diffeomorphisms
and thus to repeat these constructions.
In particular, the continuations of the hyperbolic points
$s_2,r_2,p_1,p_2,q_1$, and $q_2$ of $F_\Psi$ are defined. We will omit the
dependence on $F$ of these continuations.
As the arguments in this section are similar to those in
Sections~\ref{s.auxiliary} and \ref{s.surgery},
some of these constructions will be just sketched. We now go to the details of our
constructions.
Consider the spheres $\pi(\SS^s)$ and $F(\pi(\SS^s))$ and
denote by $\Delta_F$ the closure of the connected component of
$M\setminus \big( \pi(\SS^s) \cup F(\pi(\SS^s)) \big)$ which is
close to $\Delta$ (i.e., the component that is in the same side of
$\pi(\SS^s)$ as $\Delta$). The set $\Delta_F$ is diffeomorphic to
$\SS^2\times [0,1]$ and varies continuously with $F$ in the
$C^1$-topology. In particular, by Remark~\ref{r.time}, if $x\in \Delta_F$ and
$F^i(x)\in \Delta_F$ then $|i|\ge 9$.
Bearing in mind the definitions of
the sets ${\mathbb L}^{s,u}_{i}(F_\Psi)$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^{s,u}(F_\Psi)$
in \eqref{e.inclusions},
we define their ``continuations" ${\mathbb L}^{s,u}_{i}(F)$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^{s,u}(F)$
for $F$ close to $F_\Psi$ by
\[
\begin{split}
{\mathbb L}^u_i(F)&=\{x\in \Delta_F \, \colon \, x \in W^u(q_i,F) \,\,
\mbox{and}\,\, F^{-i}(x) \notin \Delta_F \,\, \mbox{for all}\,\, i\ge
2\},\\
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_i(F)&=\{x\in \Delta_F \, \colon \, x \in W^u(p_i,F) \,\,
\mbox{and}\,\, F^{-i}(x) \notin \Delta_F \,\, \mbox{for all}\,\, i\ge
2\},
\\
{\mathbb L}^s_i(F)&=\{x\in \Delta_F \, \colon \, x \in W^s(p_i,F) \,\,
\mbox{and}\,\, F_\Psi^{i}(x) \notin \Delta_F \,\, \mbox{for all}\,\,
i\ge 2\},
\\
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_i(F)&=\{x\in \Delta_F \, \colon \, x \in W^s(q_i,F) \,\,
\mbox{and}\,\, F^{i}(x) \notin \Delta_F \,\, \mbox{for all}\,\, i\ge
2\}.
\end{split}
\]
\begin{rema}\label{r.continuousinextremis}
{\em{The sets ${\mathbb L}^{s,u}_i(F)$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^{s,u}_i(F)$, $i=1,2$,
depend continuously on $F$.}}
\end{rema}
Note that the closed curves $\Gamma_{i,j}$ are normally hyperbolic
for $F_\Psi$ (recall Remark~\ref {r.paredesFPsi}).
Thus for every $F$ close to $F_\Psi$ there
are defined their continuations, denoted by
$\Gamma_{i,j}(F)$, that depend continuously on $F$.
These curves join the saddles $p_i$ and $q_j$ and
their interiors are center stable manifolds of $p_i$ and center
unstable manifolds of $q_j$. Finally, from the normal
hyperbolicity of $\Gamma_{i,j}(F)$, compact parts of the invariant
manifolds $W^s(\Gamma_{i,j}(F))$ and $W^u(\Gamma_{i,j}(F))$ depend
continuously on $F$.
Observe that $W^u(q_i,F)\setminus W^{uu}(q_i,F)$ (resp.
$W^s(p_i,F)\setminus W^{ss}(p_i,F)$) has two connected
components (separatrices), denoted by $W^{u,+}(q_i,F)$ and
$W^{u,-}(q_i,F)$ (resp. $W^{s,\pm}(p_i,F)$). We
choose these components such that the following holds:
\begin{rema}[Invariant manifolds of $\Gamma_{i,j}(F)$]
\label{r.compatibleintersections}
$\,${\em{
\begin{itemize}
\item
$W^u(\Gamma_{1,j}(F))\setminus W^u(p_1,F)= W^u(\Gamma_{2,j}(F))\setminus W^u(p_2,F)= W^{u,+}(q_j,F)\cup W^{uu}(q_j,F)$.
\item
$W^s(\Gamma_{i,1}(F))\setminus W^s(q_1,F) = W^s(\Gamma_{i,2}(F))\setminus W^s(q_2,F)= W^{s,+}(p_i,F)\cup$\newline
$W^{ss}(p_i,F)$.
\item $W^u(\Gamma_{i,1}(F))\cap W^u(\Gamma_{i,2}(F))= W^u(p_i,F)$.
\item $W^s(\Gamma_{1,j}(F))\cap W^s(\Gamma_{2,j}(F))= W^s(q_j,F)$.
\end{itemize}}}
\end{rema}
By Lemma~\ref{l.centerbundles} and using the notation in Remark~\ref{r.notation}, for every $F$ close
to $F_\Psi$ there is a unique invariant central bundle $E^{cs}_F$
(resp. $E^{cu}_F$)
defined on
$W^u(p_i,F)$ (resp. $W^s(q_j,F)$) and transverse to the strong stable (resp. unstable) direction at $p_i$
(resp. $q_j$).
The central bundles $E^{cs}_F$ and $E^{cu}_F$
depend continuously on $F$.
\begin{rema}\label{r.compatiblecentral}
{{\em
The manifolds with boundary $W^u(\Gamma_{i,1}(F))$ and
$W^u(\Gamma_{i,2}(F))$ are tangent along $W^u(p_i,F)$ to the plane
field $E^{cs}_F$. As in Remark~\ref{r.semiplanes}, for $x\in
W^u(p_i,F)$, the vectors of $E^{cs}_F$ entering in the interior of
$W^u(\Gamma_{i,j}(F))$ define an half plane $E^{cs}_{+,F}(x)$.
Analogously,
the manifolds with boundary $W^s(\Gamma_{1,j}(F))$ and
$W^u(\Gamma_{2,j}(F))$ are tangent along $W^s(q_j,F)$ to the plane
field $E^{cu}_F$. For $x\in W^s(p_j,F)$, the vectors of $E^{cu}_F$
entering in the interior of $W^(\Gamma_{i,j}(F))$ define a half plane
$E^{cu}_{+,F}$.
}}
\end{rema}
As in \eqref{e.Gaqi} and \eqref{e.Gapi}, we
define the sets
$$
\Gamma(q_j,F)\stackrel{\scriptscriptstyle\rm def}{=} \Gamma_{1,j}(F)\cup\Gamma_{2,j}(F) \quad \mbox{and}
\quad \Gamma(p_i,F)\stackrel{\scriptscriptstyle\rm def}{=} \Gamma_{i,1}(F)\cup\Gamma_{i,2}(F).
$$
Then
$$
W^u(\Gamma(q_j,F))= W^u(\Gamma_{1,j}(F))\cup W^u(\Gamma_{2,j}(F))
$$
is a $C^1$-surface with boundary
whose compact parts
depend continuously on $F$. Moreover,
$$
W^u(\Gamma(q_1,F))\cap W^u(\Gamma(q_2,F))= W^u(p_1,F)\cup W^u(p_2,F)
$$
The surfaces $W^u(\Gamma(q_1,F))$ and $W^u(\Gamma(q_2,F))$
depend continuously on $F$ and
are
tangent to $E^{cs}_F$ along this intersection. This last assertion is just a
version of Lemma~\ref{l.tangent} for $F$ close to $F_\Psi$.
Similarly, the
compact parts of the surfaces $W^s(\Gamma(p_1,F))$ and $W^s(\Gamma(p_2,F))$ depend
continuously on $F$ and they are tangent to $E^{cu}_F$
along their intersection $W^s(q_1,F)\cup W^s(q_2,F)$.
Using the previous notation we get that
\[
\begin{split}
&\overline{{\mathbb L}^u_j(F)}={\mathbb L}^u_j(F)\cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F)\cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2(F),\\
& \overline{{\mathbb L}^s_i(F)}={\mathbb L}^s_i(F)\cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)\cup {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_2(F).
\end{split}
\]
The set $\overline{{\mathbb L}^u_j(F)}$ is the connected
component of $W^u(\Gamma(q_j,F))\cap\Delta_F$ whose negative iterates
$F^{-i}(\overline{{\mathbb L}^u_j(F)})$, $i\ge 2$,
are disjoint from $\Delta_F$. Similarly, the set $\overline{{\mathbb L}^s_i(F)}$ is the connected
component of $W^u(\Gamma(p_i,F)\cap\Delta_F$ whose positive iterates
larger than $2$ are disjoint from $\Delta_F$.
As a consequence of the previous constructions, we get
\begin{lemm}\label{l.rectanglescontinuous}
The sets
$\overline{{\mathbb L}^u_j(F)}$ and $\overline{{\mathbb L}^s_i(F)}$ are ``rectangles"
depending continuously on $F$ (for the $C^1$ topology).
\end{lemm}
The sets $\GG^{s,u} (F)$ and ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{s,u} (F)$ are defined similarly as in the case
$F_\Psi$. The set $\GG^u(F)$ is the topological disk with two cusps
(these cuspidal points are in $W^u(p_1,F)$ and $W^u(p_2,F)$) whose boundary
is the union of $W^u(\Gamma(q_1,F))\cap \Delta_F$ and
$W^u(\Gamma(q_2,F))\cap \Delta_F$. There is an analogous definition for $\GG^s(F)$.
Note that by construction these sets depend continuously on $F$.
Finally, the set ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)$ is the topological ball bounded by $\overline{{\mathbb L}^u_1(F)}$, $\overline{{\mathbb L}^u_2(F)}$, $\GG^u(F)$ and $F(\GG^u(F))$ that is close
to ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F_\Psi)$. There is a similar definition for the set ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F)$.
By construction the sets ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{s,u}(F)$ depend continuously on $F$.
There is the following reformulation of
Corollary~\ref{c.inextremis} and Lemma~\ref{l.hetero} for
diffeomorphisms $F$ close to $F_\Psi$.
\begin{lemm}
\label{l.inextremis}
Consider a diffeomorphism $F$ close to $F_\Psi$.
\begin{enumerate}
\item \label{F1}
If $x\in \Delta_F$ is a chain recurrent point for $F$ then $x\in
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F) \cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)$.
\item \label{F2}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)\ne\emptyset$ then $F$ has a
heterodimensional cycle associated to $p_1$ and $q_1$.
\item \label{F3}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F) \pitchfork {\mathbb L}^s_1 (F)\neq \emptyset$ then the
homoclinic class of $q_1$ is non trivial.
\item \label{F4}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F) \pitchfork {\mathbb L}^u_1 (F) \neq \emptyset$ then
homoclinic class of $p_1$ is non trivial.\item
\label{F5} If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1 (F) \cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s (F) = \emptyset$ then
$p_1$ is isolated.
\item
\label{F6}
If ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1 (F) \cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F) = \emptyset$ then
$q_1$ is isolated.
\end{enumerate}
\end{lemm}
Observe also that by construction
$$
\Delta_F\subset {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F)
\subset
W^s(s_2,F)
\quad \mbox{and} \quad
\Delta_F\setminus {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)
\subset
W^u(r_2,F).
$$
As in Corollary~\ref{c.inextremis},
a consequence of these inclusions is the following.
\begin{lemm}
For every $F$ close to $F_\psi$, every point of $\Delta_F$ that is chain recurrent is contained in ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F)\cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)$.
\end{lemm}
\section{Choice of the local diffeomorphism $\Psi$} \label{s.psi}
Lemmas~\ref{l.bananasfirst} means that,
for the diffeomorphism $F_\Psi$,
the existence
of heterodimensional cycles and homoclinic
intersections for $p_1,p_2,q_1$, and $q_2$
depend on the intersections of the sets ${\mathbb L}^{u,s}_i
(F_\Psi)$, ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^{u,s}_i (F_\Psi)$, ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u (F_\Psi)$, and
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F_\Psi)$. The choice of the identification map
$\Psi$ determines these intersections. Lemma~\ref{l.inextremis}
explains how these properties are translated for diffeomorphisms
close to $F_\Psi$.
We assume that the local diffeomorphism $\Psi$ is such that
the diffeomorphism $F_\Psi$
satisfies the following two conditions, see Figure~\ref{f.td}:
\smallskip
\noindent {\bf{(T) Topological hypothesis:}} There is a point
$z\in \Delta$ such that
$$\
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F_\Psi)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)=
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F_\Psi)\cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F_\Psi)={\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)\cap
{\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F_\Psi)= \{z\}.
$$
\noindent {\bf{ (D) Differentiable hypothesis:}}
\begin{itemize}
\item
The intersection of the semi-planes $E^{cs}_{+}(z)\cap
E^{cu}_{+}(z)$ is a half straight line. This implies that the
intersection ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F_\Psi)$ at $z$ is quasi-transverse.
\item
In a neighborhood of $z$, the sets ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F_\Psi)$ and
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F_\Psi)$ are locally in the same side of any locally
defined surface containing the curves ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)$ and
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F_\Psi)$.
\end{itemize}
\begin{figure}[htb]
\psfrag{r}{$r$}
\psfrag{Y1s}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^s (F_\Psi)$}
\psfrag{Y1u}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^u (F_\Psi)$}
\psfrag{Eu}{${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{u}(F_\Psi)$}
\psfrag{Es}{${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^{s}(F_\Psi)$}
\psfrag{Ecu+}{$E^{cu}_+$}
\psfrag{Ecs+}{$E^{cs}_+$}
\includegraphics[width=5cm]{TDrev.eps}\hspace{1cm}
\caption{Conditions T and D}
\label{f.td}
\end{figure}
\begin{prop}\label{p.subsigma}
Suppose that that $F_\Psi$
satisfies conditions (T) and (D). Then there
are a $C^1$-neighborhood $\cU_{F_\Psi}$ of $F_\Psi$ and a
codimension one submanifold $\Sigma$ containing $F_\Psi$
such that:
\begin{enumerate}
\item
The set $\cU_{F_\Psi}\setminus \Sigma$ is the union of two
disjoint open sets $\cU^+_{\Sigma}$ and $\cU^-_{\Sigma}$ such
that:
\begin{itemize}
\item
for every $G\in \cU^+_{\Sigma}$ the saddle $p_{1}$ is isolated
and the homoclinic class of $q_{1}$ is non-trivial, and
\item
for every $G\in \cU^-_{\Sigma}$ the saddle $q_1$ is isolated
and the homoclinic class of $p_1$ is non-trivial.
\end{itemize}
\item
Every diffeomorphism $G\in \Sigma$ has a heterodimensional
cycle associated to $p_{1}$ and $q_{1}$.
\end{enumerate}
\end{prop}
We have the following corollary:
\begin{coro}\label{co.fragile}
The submanifold $\Sigma$ consists of diffeomorphisms $F$
having fragile cycles associated to $p_1$ and $q_1$.
\end{coro}
Note that Theorem~\ref{t.fragil} follows immediately from Proposition~\ref{p.subsigma} and Corollary~\ref{co.fragile}.
Corollary~\ref{co.fragile} is a consequence of Proposition~\ref{p.subsigma}
and the following simple fact about cycles and chain recurrence
classes:
\begin{rema}[Cycles and chain recurrence classes]\label{r.cycleschains}
{\em{Let $F$ be a diffeomorphism with a heterodimensional cycle
associated to two transitive hyperbolic sets $L$ and $K$.
Then every pair of saddles $p\in L$ and $q\in K$ are in the same chain
recurrence class of $F$. In particular, both saddles $p$ and $q$ are not isolated.}}
\end{rema}
\begin{proof}[Proof of Corollary~\ref{co.fragile}]
We argue by contradiction. If there is $F\in \Sigma$ such
that the cycle is not fragile then there is a diffeomorphism $G$ close to $F$ with a
robust cycle associated to a pair of transitive hyperbolic sets
$L\ni p_1$ and $K\ni q_1$. Then by
Remark~\ref{r.cycleschains} the saddles $p_1$ and $q_1$
are not isolated for $G$. Finally, as the cycle is robust, we can assume that $G\not\in
\Sigma$, that is, $G\in \cU_\Sigma^+\cup \cU_\Sigma^-$. Since $p_1$ is isolated if $G\in \cU^+_{\Sigma}$
this implies that $G\not\in \cU^+_{\Sigma}$. Similarly, as $q_1$
is isolated if $G\in \cU^-_{\Sigma}$ we have that $G\not\in
\cU^-_{\Sigma}$. This contradicts the fact that the cycle associated to $G$ is robust.
\end{proof}
\subsection{Proof of Proposition~\ref{p.subsigma}}
To define the submanifold $\Sigma$
we take the unitary vector $\overrightarrow n$ normal to the plane
$T_z({\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_{1} (F_\Psi))\oplus T_z({\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_{2}(F_\Psi))$ pointing
to the ``opposite" direction of ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F_\Psi)$ and
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F_\Psi)$. Fix small $\epsilon>0$ and for $F$ close to $F_\Psi$
consider the family of one-dimensional disks
$$
\left\{
{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_{1}(F)+t\, \overrightarrow{n}\right\}_{t\in [-\epsilon,
\epsilon]}.
$$
The main step of the proof of the proposition if the following lemma whose
proof we postpone.
\begin{lemm}\label{l.preprop}
There is a small
neighborhood $\cU_{F_\Psi}$ of $F_{\psi}$ such that for every $F\in \cU_{F_\Psi}$
there is a unique parameter $t=\tau_F$, depending continuously on $F$,
such that
$$
\big( {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)+\tau_F\,
\overrightarrow{n}\big) \cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2(F)\ne\emptyset.
$$
There are the following three possibilities according to the value of
$\tau_F$:
\begin{itemize}
\item If $\tau_F=0$ then the diffeomorphism $F$ has a heterodimensional
cycle associated to $p_1$ and $q_1$.
\item If $\tau_F>0$ then $p_{1}$ is isolated
and the homoclinic class of $q_1$ is non-trivial.
\item If $\tau_F<0$ then $q_{1}$ is isolated
and the homoclinic class of $p_1$ is
non-trivial.
\end{itemize}
\end{lemm}
\begin{figure}[htb]
\psfrag{z}{$z$}
\psfrag{Y1s}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^s(F)$}
\psfrag{Y1u}{${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}_{1}^u (F)$}
\psfrag{text1}{$q_1$ is isolated}
\psfrag{text2}{$H(p_1)$ is non-trivial}
\psfrag{text3}{$p_1$ is isolated}
\psfrag{text4}{$H(q_1)$ is non-trivial}
\includegraphics[width=8cm]{isolated.eps}\hspace{1cm}
\caption{Lemma~\ref{l.preprop}}
\label{f.isolated}
\end{figure}
\begin{proof}[Proof of Proposition~\ref{p.subsigma}]
In view of Lemma~\ref{l.preprop} we let
$$
\Sigma\stackrel{\scriptscriptstyle\rm def}{=} \{ F\in \cU_{F_\Psi}\, \mbox{such that $\tau_F=0$}\}.
$$
Then the set $\cU_{F_\Psi} \setminus \Sigma$
has two components $\cU_{\Sigma}^+$ and $\cU_\Sigma^-$.
The component $\cU_{\Sigma}^+$
consists of the
diffeomorphisms $F$ such that $\tau_F>0$ and
$\cU_{\Sigma}^-$ consists of the
diffeomorphisms $F$ with $\tau_F<0$. The proposition now follows immediately from
Lemma~\ref{l.preprop}.
\end{proof}
\begin{proof}
Consider the surface
$$
{\mathcal Y}} \def\cZ{{\mathcal Z}^s_1(F)=\bigcup_{t\in [-\varepsilon,\varepsilon]} {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_{1}(F)+t\, \overrightarrow{n}.
$$
Recall that ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F)$ depend
continuously on $F$ (see Remark~\ref{r.continuousinextremis}).
Thus the surface ${\mathcal Y}} \def\cZ{{\mathcal Z}^s_1(F)$ also depends
continuously on $F$. Note that, by hypothesis, ${\mathcal Y}} \def\cZ{{\mathcal Z}^s_1(F_\Psi)$
is transverse to ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F_\Psi)$ and this intersection is
the point $z$ in condition (T). Thus for $F$ close to $F_\Psi$ the
intersection ${\mathcal Y}} \def\cZ{{\mathcal Z}^s_1(F) \cap{\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_{1} (F)$ also consists of
exactly one point $z_F$ depending continuously on $F$. Thus there is
exactly one parameter $\tau_F$ with
$$
z_F \in {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F) + \tau_F\, \overrightarrow n.
$$
Moreover, the parameter
$\tau_F$
depends continuously on $F$.
If $\tau_F=0$ then ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F)\cap {\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_2(F)
\ne\emptyset$. Thus by item (\ref{F2}) in Lemma~\ref{l.inextremis}
the diffeomorphism $F$ has a heterodimensional cycles associated
to $p_1$ and $q_1$.
The choice of $\overrightarrow n$ implies that for $\tau_F>0$ the sets
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_{1}(F)$ and ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s(F)$ are disjoint. Thus
by item (\ref{F5}) in Lemma~\ref{l.inextremis}
the point
$p_{1}$ is isolated for $F$.
Similarly, for $\tau_F>0$ one has that ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)$ intersects
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)$. Thus, if $\tau_F$ is small enough, one has that
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1(F)\pitchfork
{\mathbb L}^u_1(F)$. Hence by item (\ref{F3}) in Lemma~\ref{l.inextremis}
the homoclinic class $H(q_1,F)$ is
non-trivial.
A similar argument shows that for $\tau_F<0$ the sets ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_{1}(F)$
and ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u(F)$ are disjoint. As above item (\ref{F6})
in Lemma~\ref{l.inextremis}
implies that
the point $q_1$ is isolated. Also for
$\tau_F<0$ one has that ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1(F)\pitchfork {\mathbb L}^s_1(F)$
and item (\ref{F4}) in Lemma~\ref{l.inextremis}
implies that the homoclinic class $H(p_1,F)$ is non-trivial.
This completes the proof of the lemma.
\end{proof}
\begin{rema}{\em{
Our construction can be done such that the saddles $q_2$ and $q_1$ are homoclinically related for every small $t>0$. For that it is enough to choose the
diffeomorphisms $\Psi$ such that ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_2(F_\Psi)\pitchfork {\mathbb L}^u_1(F_\Psi)$,
see Corollary~\ref{c.bananas}.
Observe that in this case one has $H(q_2,F_t)=H(q_1,F_t)$ for $t>0$.
However, for $t=0$ the saddle $q_1$ escapes from the non-trivial homoclinic
class of $q_2$. Surprisingly, in such a case one can generate robust cycles associated to $q_2$ and $p_1$ (this follows from \cite{BDK}) but not associated to $q_1$ and $p_1$.
}}
\label{r.recall}
\end{rema}
\section{Discussion} \label{s.conclusion}
Our construction provides examples of fragile cycles relating two
saddles $p_1$ and $q_1$ of different indices (for simplicity we will
omit the dependence on the diffeomorphisms). This
construction has a ``prescribed" part concerning the
relative positions of the invariant manifolds of these saddles. But
this prescribed dynamics involves only the ``cuspidal" regions
${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^s_1$ and ${\mathbb Y}} \def\mbox{Z\hspace{ -.3em}Z}{{\mathbb Z}^u_1$ of
${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s$ and ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u$, respectively. The rest of the intersection ${\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^s\cap {\mathbb E}} \def\mbox{I\hspace{ -.15em}F}{{\mathbb F}} \def\GG{{\mathbb G}^u$ can
be chosen arbitrarily. This provides a
lot of ``freedom"
for the global dynamics, for instance, for the
behavior of the other
saddles of the diffeomorphism and their invariant manifolds
(recall Remark~\ref{r.recall}).
Hence, without further assumptions on the dynamics outside the
cycle the global dynamics cannot be described.
\subsection{Partial hyperbolicity, wild dynamics, and
fragile cycles}
A first ingredient of our construction is the
gluing map $\Psi$ that plays a key role for determining the
resulting global dynamics.
\begin{ques}\label{q.ph}
Can the resulting dynamics be partially hyperbolic? More
precisely, does it exist a gluing map $\Psi$ so that
the chain recurrence class of $p_1$ for $F_\Psi$ has a partially hyperbolic
splitting with three $1$-dimensional bundles?
\end{ques}
We expect a positive answer to this question. This would
imply that the phenomena associated to these fragile cycles
would also occur in the most rigid non-hyperbolic setting of partial
hyperbolicity with one-dimensional center.
\smallskip
We next discuss how fragile cycles can be involved in
the generation of {\emph{wild dynamics}}
(roughly, persistent coexistence of infinitely
many homoclinic classes, see \cite{bible} for further details).
First note that perturbations of the diffeomorphism $F$ in
the fragile cycles submanifold
$\Sigma$ in Theorem~\ref{t.fragil} expels periodic
points in the chain recurrence class that
simultaneously
contains $p_1$ and $q_1$:
the saddle $p_1$ is expelled by diffeomorphisms in the component $\cU^+_{\Sigma}$ and the saddle $q_1$ by diffeomorphisms in the component $\cU^-_{\Sigma}$.
Thus the chain recurrence class fall into pieces.
On the other hand,
there is a natural question about whether fragile cycles may generate a ``cascade" of fragile cycles:
\begin{ques}\label{q.repeat}
Can the submanifold of fragile cycles $\Sigma$ be accumulated by (co-dimension one) submanifolds consisting of fragile cycles?
\end{ques}
If the answer to this question would be positive then a ``fragile cycles
configuration" could be repeated generating infinitely many different
chain recurrence classes.
Being very optimistic, positive answers to the two questions above
could provide the first examples of
wild dynamics in the partially hyperbolic (with
one-dimensional center) setting.
We will discuss further questions about wild dynamics later.
\subsection{Chain recurrence classes}\label{ss.chain}
Consider a diffeomorphism $F$ in the submanifold $\Sigma$ of fragile
cycles. By construction, the intersection
$W^s(p_1)\pitchfork W^u(q_1)$ contains a curve $\Gamma_{1,1}$ joining
$p_1$ to $q_1$. Moreover, the intersection $W^u(p_1)\cap W^u(q_1)$
is exactly the orbit of a quasi-transverse heteroclinic point
$x_F$. This cyclic configuration implies that the chain recurrence
classes of $p_1$ and $q_1$ coincide and contain the curve $\Gamma_{1,1}$
and the orbit of $x_F$,
recall also Remark~\ref{r.cycleschains}.
We can now perform
perturbations of $F\in \Sigma$ preserving the cycle, that is, the
resulting diffeomorphisms continue to belong to $\Sigma$. In this way, and using
for instance
the arguments in \cite{BD08}, one can slightly modify the central
eigenvalues of the saddles $p_1$ and $q_1$ in the cycle
(corresponding to the tangent
direction of the connection $\Gamma_{1,1}$)
to generate ``new" periodic saddles $r$ whose
orbits pass arbitrarily close to $p_1$ and $q_1$ and belong
to the chain recurrence class of $C(p_1)=C(q_1)$. The latter fact is a consequence of the geometry of the cycle that guarantees that indeed $W^u(r)\cap W^s(p_1)\ne \emptyset$ and $W^s(r)\cap W^u(q_1)\ne\emptyset$, see Figure~\ref{f.twisted}.
\begin{figure}[htb]
\psfrag{x}{$x$}
\psfrag{y}{$y$}
\psfrag{r}{$r$}
\psfrag{p}{$p_1$}
\psfrag{q}{$q_1$}
\psfrag{g11}{$\Gamma_{1,1}$}
\includegraphics[width=4.2cm]{twistedrev.eps}\hspace{1cm}
\caption{Periodic points in $C(p_1)=C(q_1)$.}
\label{f.twisted}
\end{figure}
\begin{ques}\label{q.Si} For $C^1$-generic diffeomorphisms
$F\in\Sigma$, is the heteroclinic curve $\Gamma_{1,1}$ contained in the
closure of a set of periodic points?
Can the periodic points in that set be chosen in the same
homoclinic class or (even more) homoclinically related?
\end{ques}
Concerning this question, observe that our construction
provides saddles that are in the same recurrence class (the one of
$p_1$ and $q_1$). Note also that by \cite{BC04}, for
$C^1$-generic diffeomorphisms, the chain recurrence classes of
periodic points always coincide with their homoclinic classes. The
difficulty here is that we consider $C^1$-generic diffeomorphisms
in the codimension one submanifold $\Sigma$ (which is a meager set).
Thus the result in \cite{BC04}
cannot be applied. Indeed, for $F\in \Sigma$ the homoclinic classes of
$p_1$ and $q_1$ are both trivial and
hence different from their non-trivial chain
recurrence classes since $p_1,q_1\in C(p_1)=C(q_1)$.
We believe that the answer to Question~\ref{q.Si} is positive.
In such a case, it would be interesting to consider the central Lyapunov exponents of the periodic orbits in the chain recurrence class $C(p_1)=C(q_1)$.
This is our next topic.
\subsection{Lyapunov exponents} The papers
\cite{DHRS09,LOR,DG} consider examples of diffeomorphisms having twisted
heterodimensional cycles (recall Section~\ref{s.introduction}) and analyze the spectrum of central Lyapunov exponents of the (non-trivial) homoclinic classes involved in the cycle. In the examples in \cite{LOR,DG} this spectrum has a {\emph{gap}}.
The existence of this gap is related to the fact that the homoclinic
class $H(p)$ considered contains a saddle $q$ of different index
that looks like a ``cuspidal corner point" (exactly as the points $p_1$ and
$q_1$ of our construction)
and satisfies $q\in H(p)$ but $H(q)=\{q\}$.
So the saddle $q$ is topologically an extreme point but also dynamically is also an extreme point as it is not homoclinically related to other saddles in the class $H(p)$.
We believe that precisely this is reflected by the Lyapunov
spectrum that has one gap.
The previous comments and the fact that
the chain recurrence classes considered in this paper have ``two cuspidal corner
points" (the saddles $p_1$ and $q_1$) that are also dynamically extremal lead to the following question.
\begin{ques} Does there exist a
gluing map $\Psi$ such that there is a neighborhood $\cU$ of
$F_\Psi$ such that ``persistently" in $\cU\cap \Sigma$ the
diffeomorphisms have two (or more) gaps in the spectrum of central Lyapunov
exponents of the periodic orbits in the chain recurrence class $C(p_1)$?
Here the term ``persistent" is purposely vague and may mean $C^1$-generic or $C^1$-dense, for instance.
\end{ques}
Observe that by \cite{ABCDW07} for a $C^1$-generic diffeomorphism the spectrum of Lyapunov exponents
of homoclinic classes has no gaps.
Hence, one would need to consider diffeomorphisms of non-generic type.
\subsection{Collision, collapse, and birth of classes}
Consider an arc of diffeomorphisms $(F_t)_{t\in[-1,1]}$ that intersects $\Sigma$ transversely at $t=0$. Assume that $F_t\in \cU^+_\Sigma$ for $t>0$ and
$F_t\in \cU^-_\Sigma$ for $t<0$.
Our construction implies that for $t>0$ the class $H(q_1,F_t)$ is non-trivial,
$H(p_1,F_t)=\{p_1\}$, and both classes are disjoint. Similarly, for $t<0$ the class $H(p_1,F_t)$ is non-trivial, $H(q_1,F_t)=\{q_1\}$, and both classes are disjoint. Since $F_0\in \Sigma$ and $H(q_1,F_0)=\{q_1\}$ and
$H(p_1,F_0)=\{p_1\}$, each of these classes collapses to a single point at $t=0$. Similarly (or symmetrically), for $t=0$ the chain recurrence classes of $p_1$ and $q_1$ collide at $t=0$ and contain the heteroclinic segment $\Gamma_{1,1}$, when one of these classes collapses to a point for $t\neq 0$. This illustrates the lower semi-continuous dependence of homoclinic classes and the upper semi-continuity of chain recurrence classes.
Let us observe that \cite{BCGP} provides a locally $C^1$-dense set of diffeomorphisms where homoclinic classes are properly contained
in a robustly isolated chain recurrence class. This construction is somewhat
similar to the one in this paper
and involves a heterodimensional cycle relating cuspidal corner points of the class.
Coming back to our construction, it would be interesting to understand how for $t>0$ the points of $H(q_1,F_t)$ escape from the class or ``disappear" as $t\to0^+$. As we discussed above, in some cases the saddle $q_1$ is accumulated by saddles of the same index when $t=0$. These saddles are not homoclinically related to $q_1$ but they are in its chain recurrence class. This indicates that there are (infinitely many) saddles that ``escape" from the homoclinic class of $q_1$ but not from the chain recurrence class of $q_1$ as $t$ evolves. Though we do not know how these saddles are homoclinically related.
Note that there is completely symmetric scenery for the saddle $p_1$.
\begin{ques}
Is there a diffeomorphism $F\in\Sigma$ with infinitely many different
homoclinic classes, all of them contained in the chain recurrence class of $p_1$ and $q_1$?
\end{ques}
In view of the previous discussion, a simpler question is if the diffeomorphisms in $\Sigma$ can be chosen having wild dynamics close to the cycle. More precisely,
\begin{ques}
Does there exist $F\in\Sigma$ with infinitely many different chain recurrence classes accumulating to $p_1$ or to $q_1$?
\end{ques}
Finally, let us observe that the previous setting is somehow reminiscent to the setting of the geometrical Lorenz attractor: at the point of ``bifurcation", when a singular cycle occurs, infinitely many
orbits of the vector field transform into heteroclinic orbits of a singularity, see for
instance \cite{BLMP}.
\section*{Acknowledgments}
This paper was partially supported by CNPq, FAPERJ, and Pronex (Brazil),
Agreement France-Brazil in Mathematics, and ANR Project DynNonHyp
BLAN08-2$_-$313375.
LJD thanks the warm hospitality of Institut de Math\`ematiques de Bourgogne.
|
\section{Introduction}
\label{sec:intro}
Cataclysmic variable stars (CVs) are a class of interacting binary system undergoing mass transfer from a Roche-lobe
filling secondary to a white dwarf primary, usually via a gas stream and accretion disc. A bright spot is formed where
the gas stream collides with the edge of the accretion disc, often resulting in an \textquoteleft orbital
hump\textquoteright~in the light curve at phases $0.6-1.0$ due to the area of enhanced emission rotating into our line
of sight. For an excellent overview of CVs, see \citet{warner1995a} and \citet{hellier2001}. The light curves of eclipsing
CVs can be quite complex, with the accretion disc, white dwarf and bright spot all being eclipsed in rapid succession.
When observed with time resolutions of the order of a few seconds, this eclipse structure allows the system parameters
to be determined to a high degree of precision with relatively few assumptions \citep{wood1986}. Over the last eight
years our group has used the high-speed, three-colour camera {\sc ultracam} \citep{dhillon2007} to obtain such
time-resolution. The ability to image in three different wave-bands simultaneously makes {\sc ultracam} an ideal tool to study the
complex, highly variable light curves of CVs. Using {\sc ultracam} data, we have obtained system parameters for several short
period systems (e.g. Feline et al. 2004a, 2004b; Littlefair et al. 2006a, 2007, 2008), \nocite{feline2004a, feline2004a,
feline2004b, littlefair2006a, littlefair2007, littlefair2008} including the first system accreting from a sub-stellar donor
(Littlefair et al. 2006b). \nocite{littlefair2006b}
Despite extensive study over recent decades, there are still several outstanding issues with evolutionary theories
of CVs that have wide ranging implications for all close binary systems. The secular evolution of CVs is driven by
angular momentum losses from the binary orbit. In the standard model, systems with orbital periods below $\sim$130
minutes are thought to lose angular momentum via gravitational radiation. Angular momentum losses sustain mass transfer
and subsequently drive the system to shorter orbital periods, until the point where the donor star becomes degenerate
\citep[e.g.][]{paczynski1981}. Here, the donor star is driven out of thermal equilibrium and begins to expand in response
to mass-loss, driving the system to longer orbital periods. We therefore expect to observe a period cut-off around
$P_{orb} \simeq 65-70$ minutes, dubbed the ``period minimum'', in addition to a build up of systems at this minimum period
(the ``period spike''; Kolb \& Baraffe 1999).\nocite{kolb1999} The period spike has recently been identified by
\citet{gaensicke2009}, whose
study of SDSS CVs found an accumulation of systems with orbital periods between 80 and 86 minutes. This is significantly
longer than expected. A larger than expected orbital period implies that the orbital separation is larger than expected,
nd thus the radii of the donor star must also be larger than expected in order to remain Roche-lobe filling. Recent
observations by \citet{littlefair2008} support this, suggesting that the donor stars in short period CVs are roughly
10 percent larger than predicted by the models of \citet{kolb1999}. The reason why the donor stars appear oversized
remains uncertain. Possible explanations include some form of enhanced angular momentum loss \citep[e.g.][]{patterson1998,
kolb1999, willems2005} which would increase mass loss and drive the donor stars further from thermal equilibrium,
or missing stellar physics in the form of magnetic activity coupled with the effects of rapid rotation \citep[e.g.][]{chabrier2007}.
One way to determine why the donor stars appear oversized is to compare the shape of the observed donor mass-period
relationship ($M_{2}-P_{orb}$), and by implication, mass-radius ($M_{2}-R_{2}$) relationship, to the models of
\citet{kolb1999}, calculated with enhanced angular momentum loss or modified stellar physics. These models, in principle,
make different predictions for the shape of the mass-period relationship and the position of the period minimum. Both the
shape of the mass-period relationship and the position of the period minimum are dependent on the ratio
$\kappa = \tau_{M}/\tau_{KH}$, where $\tau_{M}$ and $\tau_{KH}$ are the mass-loss and thermal timescales of the donor star
respectively. Initially, $\kappa$$\gg$1, and the donor is able to contract in response to mass-loss. As the system evolves
to shorter orbital periods both timescales increase, although the thermal timescale increases much faster than the mass-loss
timescale. This results in $\kappa$ decreasing with orbital period. When the two timescales become comparable, the donor
is unable to contract rapidly enough to maintain thermal equilibrium and becomes oversized for a given mass. Since donor
expansion does not occur until the thermal and mass-loss timescales become comparable, if enhanced angular momentum loss
is responsible for the oversized CV donors, the systems {\it immediately} below the period gap would not be expected to be
far from thermal equilibrium. In contrast, star spots would inhibit the convective processes in all CV donors below the
period gap (assuming of course that spot properties are similar at all masses). Models that include the effects of enhanced
angular momentum loss and star spot coverage therefore begin to diverge significantly in $\dot{M}$ and $M_{2}$ at orbital
periods of 100 minutes. We can distinguish between these models if we have a sample of CVs that covers a wide range of
orbital periods and whose component masses and radii are known to a high degree of precision (e.g. $\sigma$$M_{2}$$\sim0.005M_{\odot}$).
Unfortunately we lack enough precise mass-radii determinations for systems with orbital periods between 95 and 130 minutes.
To overcome this shortage we observed eclipses of three CVs below the period gap: CTCV J1300-3052, CTCV J2354-4700 and SDSS
J115207.00+404947.8 (hereafter CTCV 1300, CTCV 2354 and SDSS 1152).
CTCV 2354 and CTCV 1300 were discovered as part of the Cal\'{a}n-Tololo Survey follow up \citep{tappert2004}. During the
follow up, both systems were found to be eclipsing with orbital periods of 94.4 and 128.1 minutes, respectively.
Basic, non-time resolved, spectroscopic data was obtained for each system. The spectrum of CTCV 1300 showed features typical
of the three main components in CVs: strong emission lines from the accretion disc, broad, shallow absorption features from
the white dwarf and red continuum and absorption bands from the donor. CTCV 2354 was found to contain strong emission lines
of H and He, generally typical of a dwarf nova in quiescence.
SDSS 1152 was identified as a CV by \citet{szkody2007}. The system shows broad, double-peaked, Balmer emission
lines, which are characteristic of a high-inclination accreting binary. Follow up work by \citet{southworth2010} found the system
to have an orbital period of 97.5 minutes.
In this paper we present {\sc ultracam} light curves of CTCV 1300 ($u'g'r'i'$), CTCV 2354 ($u'g'r'$) and SDSS 1152 ($u'g'r'$),
and in each case attempt to determine the system parameters via light curve modelling. In addition, we also present an updated
analysis of all eclipsing systems previously published by our group: OU Vir (Feline et al. 2004a),
\nocite{feline2004a} XZ Eri and DV UMa (Feline et al. 2004b), \nocite{feline2004b} SDSS J1702+3229 (Littlefair et al. 2006a),
\nocite{littlefair2006a} SDSS J1035+0551 (Littlefair et al. 2006b, 2008), \nocite{littlefair2006b, littlefair2008}
SDSS J150722+523039 \citep{littlefair2007, littlefair2008}, SDSS J0903+3300, SDSS J1227+5139, SDSS J1433+1011, SDSS J1501+5501
and SDSS J1502+3334 \citep{littlefair2008}. Our primary reason for doing so was the introduction of a new analysis utilising
Markov Chain Monte Carlo (MCMC) techniques and an updated bright spot model. The MCMC analysis is more reliable at converging to a best
fit than the downhill simplex algorithm used previously, while the new bright spot model allows for a more realistic modelling of
bright spot dominated systems (e.g. CTCV 1300, DV UMa, SDSS 1702) and should thus provide more accurate values of the mass ratio, $q$.
While implementing these changes, we also discovered a bug in the code previously used to bin the light curves (see e.g.
Littlefair et al. 2006a \nocite{littlefair2006a} for details of the original code) which resulted in the white
dwarf radius being underestimated, and consequently, the white dwarf and donor mass being overestimated. Full details
are provided in section \ref{sec:pme}.
\section{Observations}
\label{sec:obs}
In Table \ref{table:journal} we present details of the observations used to analyse CTCV 1300, CTCV 2354, SDSS 1152 and
SDSS 1501. For observations
of other systems we refer the reader to table 1 in the following publications: Feline et al. 2004a (OU Vir), Feline et al.
2004b (XZ Eri and DV UMa), Littlefair et al. 2006a (SDSS 1702), Littlefair et al. 2007 (SDSS 1507) and Littlefair et al. 2008
(SDSS 0903, SDSS 1227, SDSS 1433, SDSS 1501 and SDSS 1502).
For reasons outlined in section \ref{sec:pme} we do not use the SDSS 1501 data listed in Littlefair et al. (2008). Instead we model
a single eclipse observed in 2004 (Table \ref{table:journal}). Not all of the eclipses listed in Table \ref{table:journal}
are used for determining system parameters. This is because the eclipses have poor signal-to-noise, or lack clear bright spot
features. The eclipses not used for determining system parameters are however still used to refine
our orbital ephemerides (section
\ref{sec:ephem}). These eclipses include CTCV 2354 cycle numbers 11197, 11198, 11366, 11396, 11457, 11472 and SDSS 1501 cycles
24718 and 24719. CTCV 2354 cycle numbers 16156, 16676 and CTCV 1300 cycle number 12888 are analysed separately in section
\ref{sec:notes} because the shape of the eclipse has changed significantly in comparison to the 2007 data (see section \ref{sec:lcm}).
Data reduction was carried out in a standard manner using the {\sc ultracam} pipeline reduction software, as described in
\citet{feline2005} and \citet{dhillon2007}. A nearby comparison star was used to correct the data for transparency variations.
Observations of the standard stars G162-66, G27-45 and G93-48 were used to correct the magnitudes to the standard SDSS system
\citep{smith2002}. Due to time constraints and poor weather, we were unable to observe a standard star to flux calibrate our
data for SDSS 1152. Consequently, we have used the Sloan magnitudes of the comparison stars and corrected for different
instrumental response. To do this, we use measured response curves for filters and dichroics to create overall response
curves for {\sc ultracam}. These are then combined with curves of theoretical extinction and library spectra
\citep{pickles1998} to obtain synthetic {\sc ultracam} colours. The same process is then repeated for the SDSS colour set,
with the difference between the two sets being the correction applied.
\begin{table*}
\caption{Journal of observations. The dead-time between exposures was 0.025~s for all observations.
The relative GPS time stamping on each data point is accurate to 50 $\mu$s. Instr setup denotes the telescope
(WHT, NTT or VLT) and instrument used for each observation, where UCAM and USPEC represent {\sc ultracam} and
{\sc ultraspec}, respectively. Phase Cov corresponds to the phase coverage of the eclipse, taking the eclipse of the
white dwarf as phase 1. $T_{exp}$ and $N_{exp}$ denote the exposure time, and number of exposures, respectively. }
\centering
\begin{tabular}{ccccrccrcc}
\hline
\hline
Date & Object & Instr setup & $T_{mid}$ (HMDJ) & Cycle & Phase Cov & Filters & $T_{exp}$ (s) & $N_{exp}$ & Seeing ('')\\
\hline
2007 June 09 & CTCV 2354 & VLT+UCAM & 54261.383926(25) & 0 & 0.73--1.08 & $u'g'r'$ & 2.22 & 821 & 0.6--1.0\\
2007 June 13 & CTCV 2354 & VLT+UCAM & 54265.316786(61) & 60 & 0.70--1.11 & $u'g'r'$ & 4.92 & 473 & 0.8--1.2\\
2007 June 15 & CTCV 2354 & VLT+UCAM & 54267.348921(21) & 91 & 0.78--1.08 & $u'g'r'$ & 2.22 & 779 & 0.6--0.7\\
2007 June 15 & CTCV 2354 & VLT+UCAM & 54267.414476(20) & 92 & 0.74--1.05 & $u'g'r'$ & 2.22 & 757 & 0.6--0.7\\
2007 June 16 & CTCV 2354 & VLT+UCAM & 54268.397717(20) & 107 & 0.72--1.06 & $u'g'r'$ & 2.22 & 845 & 0.6--1.1\\
2007 June 19 & CTCV 2354 & VLT+UCAM & 54271.413077(21) & 153 & 0.82--1.15 & $u'g'r'$ & 2.22 & 826 & 0.6--1.0\\
2007 June 20 & CTCV 2354 & VLT+UCAM & 54272.396368(29) & 168 & 0.50--1.50 & $u'g'r'$ & 2.32 & 2390 & 0.6--1.0\\
2007 June 21 & CTCV 2354 & VLT+UCAM & 54273.314054(5) & 182 & 0.86--1.20 & $u'g'r'$ & 1.96 & 931 & 1.2--2.4\\
2007 June 21 & CTCV 2354 & VLT+UCAM & 54273.379579(3) & 183 & 0.77--1.09 & $u'g'r'$ & 1.96 & 916 & 0.9--1.5\\
2009 June 12 & CTCV 2354 & NTT+USPEC & 54995.350263(6) & 11197 & 0.65--1.35 & $ g' $ & 9.87 & 482 & 1.4--2.6\\
2009 June 12 & CTCV 2354 & NTT+USPEC & 54995.415961(6) & 11198 & 0.35--1.17 & $ g' $ & 9.87 & 482 & 1.0--2.2\\
2009 June 23 & CTCV 2354 & NTT+USPEC & 55006.428224(2) & 11366 & 0.66--1.16 & $ g' $ & 3.36 & 817 & 1.4--2.2\\
2009 June 25 & CTCV 2354 & NTT+USPEC & 55008.394766(1) & 11396 & 0.70--1.22 & $ g' $ & 3.36 & 855 & 1.2--2.0\\
2009 June 29 & CTCV 2354 & NTT+USPEC & 55012.393334(1) & 11457 & 0.55--1.37 & $ g' $ & 2.98 & 1593 & 0.8--2.4\\
2009 June 30 & CTCV 2354 & NTT+USPEC & 55013.376588(1) & 11472 & 0.30--1.55 & $ g' $ & 1.96 & 3592 & 1.2--2.0\\
2010 May~ 03 & CTCV 2354 & NTT+UCAM & 55320.414(2) & 16156 & 0.43--1.21 & $u'g'r'$ & 8.23 & 526 & 1.4--1.8\\
2010 June 06 & CTCV 2354 & NTT+UCAM & 55354.434846(24) & 16676 & 0.47--1.19 & $u'g'r'$ & 3.84 & 1048 & 1.0--1.2\\
\hline
2007 June 10 & CTCV 1300 & VLT+UCAM & 54262.099145(3) & 0 & 0.72--1.20 & $u'g'r'$ & 1.00 & 3462 & 0.6--1.2\\
2007 June 13 & CTCV 1300 & VLT+UCAM & 54262.123093(8) & 34 & 0.74--1.15 & $u'g'i'$ & 1.95 & 1573 & 0.6--1.1\\
2010 June 07 & CTCV 1300 & NTT+UCAM & 55355.002677(1) & 12288 & 0.85--1.12 & $u'g'r'$ & 2.70 & 511 & 0.8--1.1\\
\hline
2010 Jan~ 07 & SDSS 1152 & WHT+UCAM & 55204.101282(9) & 0 & 0.16--1.13 & $u'g'r'$ & 3.80 & 1492 & 2.0--3.8\\
2010 Jan~ 07 & SDSS 1152 & WHT+UCAM & 55204.169031(8) & 1 & 0.72--1.12 & $u'g'r'$ & 3.80 & 600 & 1.4--2.5\\
2010 Jan~ 07 & SDSS 1152 & WHT+UCAM & 55204.236742(7) & 2 & 0.85--1.12 & $u'g'r'$ & 3.80 & 415 & 1.2--3.2\\
\hline
2004 May~ 17 & SDSS 1501 & WHT+UCAM & 53142.921635(6) &-11546 & 0.80--1.21 & $u'g'r'$ & 6.11 & 335 & 1.0--1.6\\
2010 Jan~ 07 & SDSS 1501 & WHT+UCAM & 55204.213149(3) & 24718 & 0.78--1.12 & $u'g'r'$ & 3.97 & 435 & 1.4--4.0\\
2010 Jan~ 07 & SDSS 1501 & WHT+UCAM & 55204.270000(3) & 24719 & 0.88--1.13 & $u'g'r'$ & 3.97 & 321 & 1.4--3.0\\
\hline
\hline
\end{tabular}
\label{table:journal}
\end{table*}
\section{Results}
\subsection{Orbital ephemerides}
\label{sec:ephem}
The times of white dwarf mid-ingress $T_{wi}$ and mid-egress $T_{we}$ were determined by locating the minimum and
maximum times, respectively, of the smoothed light-curve derivative. Mid-eclipse times, $T_{mid}$, were determined
by assuming the white dwarf eclipse to be symmetric around phase one and taking $T_{mid}=(T_{wi}+T_{we})/2$. Eclipse
times were taken from the literature for CTCV 1300, CTCV 2354 \citep{tappert2004}, SDSS 1501 \citep{littlefair2008}
and SDSS 1152 \citep{southworth2010} and combined with our mid-eclipse times shown in Table \ref{table:journal}. The
errors on our data were adjusted to give $\chi^{2} = 1$ with respect to a linear fit. In each case we observe no cycle
ambiguity. We do however, observe a significant, O-C offset between our data and the times published by \citet{tappert2004}
for CTCV 1300 and CTCV 2354. For CTCV 1300, the average difference is 165.9 seconds, while for CTCV 2354 it is
148.0 seconds. We believe this to be due to the differing methods of calculating $T_{mid}$; Tappert et al. (2004) calculated
$T_{mid}$ by fitting a parabola to the overall eclipse structure, whereas we determined $T_{mid}$ from the white dwarf
eclipse. We therefore subtract these average offsets from the published literature values and take the resulting O-C
difference as our uncertainty on that time. Where possible, we averaged the measured mid-eclipse times in the $r'$ and
$g'$ bands to fit the ephemeris. Due to the low signal-to-noise of the eclipse of CTCV 2354 in May 2010, measuring the
times of white dwarf mid-ingress $T_{wi}$ and mid-egress $T_{we}$ was not possible. Consequently, the eclipse times and
errors were measured by eye in the $g'$ and $r'$ bands and then averaged in order to estimate the cycle
number. This time was not used to refine the ephemeris but is included for completeness. The ephemerides found are shown
in Table \ref{table:ephemeris}.
\begin{table}
\caption{Orbital ephemerides.}
\centering
\begin{tabular}{ccc}
\hline
Object & $T_{\rmn{0}}$ (HJD) & $P_{\rmn{orb}}$ (d)\\
\hline
CTCV 1300 & 2454262.599146 (8) & 0.088940717 (1) \\
CTCV 2354 & 2454261.883885 (5) & 0.065550270 (1) \\
SDSS 1152 & 2455204.601298 (6) & 0.067721356 (3) \\
SDSS 1501 & 2453799.710832 (3) & 0.0568412623(2) \\
\hline
\end{tabular}
\label{table:ephemeris}
\end{table}
\subsection{Light curve morphology and variations}
\label{sec:lcm}
\subsubsection{CTCV J1300-3052}
Fig. \ref{figure:lightcurves} (top) shows the two observed eclipses of CTCV 1300 from the 2007 data
set folded on orbital phase in the $g'$ band. The white dwarf ingress and egress features are clearly
visible at phases 0.965 and 1.040, respectively, as are the bright spot features at phases 0.960 and
1.085. These features dominate the light curve, which follows a
typical dwarf nova eclipse shape (e.g. Littlefair et al. 2006a, 2007, 2008). The depth of the
bright spot eclipse indicates that the bright spot is the dominant source of light in this system,
while the eclipse of the accretion disc is difficult to discern by eye, indicating that the accretion
disc contributes little light to this system. Closer inspection of the eclipses from each night reveals
a noticeable difference in the shape of the bright spot ingress feature. This is caused by heavy pre-eclipse
flickering, and is clearly visible in Fig. \ref{fig:eclipses}. The flickering is reduced between
phases corresponding to the white dwarf ingress and bright spot egress, indicating the source of
the flickering is the inner disc and/or bright spot. Due to the heavy flickering, we decided to fit
each night individually rather than fit to a phase-folded average, in order to provide a more robust
estimation of our uncertainties. Our 2010 observations are discussed in section \ref{sec:ctcv1300_2010}.
\subsubsection{CTCV J2354-4700}
\label{sec:2354}
Fig. \ref{figure:lightcurves} (middle) shows all of the observed eclipses from the 2007 data set of
CTCV 2354 folded on orbital phase in the $g'$ band. The white dwarf ingress and egress features are
clearly visible and along with the accretion disc dominate the shape of the light curve. A weak bright
spot ingress feature is visible at an orbital phase of 0.995, however the system suffers from heavy
flickering, making it difficult to identify the bright spot egress. The shape of the average light curve
indicates possible egress features at phases 1.060 and 1.080, but given the scatter we cannot be certain
whether these represent genuine egress features or merely heavy flickering. The flickering is reduced between
phases corresponding to the white dwarf ingress and egress, indicating the source of the flickering is
the inner disc. Our observations from 2009 and 2010 are discussed in section \ref{sec:ctcv2354_2010}.
\subsubsection{SDSS J1152+4049}
Fig. \ref{figure:lightcurves} (bottom) shows all of the observed eclipses from the 2010 data set of SDSS 1152
folded on orbital phase in the $g'$ band. The signal-to-noise ratio of our data is low in comparison to other
systems, but we still see a clear bright spot ingress feature at phase 0.975 in addition to a clear bright spot
egress feature at phase 1.075. The white dwarf features are clear, and dominate the overall shape of the light
curve. Like CTCV 1300, the eclipse of the accretion disc is difficult to discern by eye, which again suggests
that the accretion disc contributes little light to this system.
\begin{figure}
\centering
\includegraphics[scale=0.40,angle=0,trim=0 0 0 0,clip]{fig1_cvs.eps}
\caption{{\sc ultracam} $g'$ band light curves of CTCV 1300 (2007, top), CTCV 2354 (2007, middle) and SDSS 1152 (2010, bottom).}
\label{figure:lightcurves}
\end{figure}
\subsection{Light curve modelling}
\label{sec:pme}
To determine the system parameters we used a physical model of the binary system to calculate eclipse light curves
for the white dwarf, bright spot, accretion disc and donor. Feline et al. (2004b) \nocite{feline2004b} showed that
this method gives a more robust determination of the system parameters in the presence of flickering than the derivative
method of \citet{wood1986}. The model itself is based on the techniques developed by \citet{wood1985} and
\citet{horne1994}, and is an adapted version of the one used by \citet{littlefair2008}. This model
relies on three critical assumptions: the bright spot lies on the ballistic trajectory from the donor
star, the donor fills its Roche lobe, and the white dwarf is accurately described by a theoretical mass-radius
relation. Obviously these assumptions cannot be tested directly, but it has been shown that the
masses derived with this model are consistent with other methods commonly employed in CVs over
a range of orbital periods \citep[e.g.][]{feline2004b, tulloch2009, copperwheat2010}.
The model used by \citet{littlefair2008} had to be adapted due to the prominent
bright spot observed in CTCV 1300 (Fig. \ref{figure:lightcurves}). The \textquoteleft old\textquoteright~model
fails to correctly model the bright spot ingress and egress features in this system satisfactorily and results in
a poor fit. We have thus adapted the model to account for a more complex bright spot by adding four new parameters,
following \citet{copperwheat2010}, bringing the total number of variables to 14. These are:
\begin{enumerate}
\item The mass ratio, $q=M_r/M_w$.
\item The white dwarf eclipse phase full-width at half-depth, $\Delta\phi$.
\item The outer disc radius, $R_{d}/a$, where $a$ is the binary separation.
\item The white dwarf limb-darkening coefficient, $U_{w}$.
\item The white dwarf radius, $R_{w}/a$.
\item The bright-spot scale, $S/a$. The bright spot is modelled as two
linear strips passing through the intersection of the gas stream and
disc. One strip is isotropic, while the other beams in a given direction.
Both strips occupy the same physical space. The intensity distribution is
given by $(X/S)^{Y}e^{-(X/S)^{Z}}$, where $X$ is the distance along the strips.
The non-isotropic strip does not beam perpendicular to its surface. Instead
the beaming direction is defined by two angles, $\theta_{tilt}$ and
$\theta_{yaw}$.
\item The first exponent, $Y$, of the bright spot intensity distribution.
\item The second exponent, $Z$, of the bright spot intensity distribution.
\item The bright-spot angle, $\theta_{az}$, measured relative to the line
joining the white dwarf and the secondary star. This allows adjustment of
the phase of the orbital hump.
\item The tilt angle, $\theta_{tilt}$, that defines the beaming direction of
the non-isotropic strip. This angle is measured out of the plane of
the disc, such that $\theta_{tilt} = 0$ would beam light perpendicular to the
plane of the disc.
\item The yaw angle, $\theta_{yaw}$. This angle also defines the beaming direction
of the non-isotropic strip, but in the plane of the disc and with respect to the
first strip.
\item The fraction of bright spot light that is isotropic, $f_{iso}$.
\item The disc exponent, $b$, describing the power law of the radial
intensity distribution of the disc.
\item A phase offset, $\phi_{0}$.
\end{enumerate}
The data are not good enough to determine the white dwarf limb-darkening coefficient, $U_{w}$,
accurately. To find an appropriate limb-darkening coefficient, we follow the procedure outlined in
\citet{littlefair2007}, whereby an estimate of the white dwarf effective temperature and mass is
obtained from a first iteration of the fitting process outlined below, assuming a limb-darkening
coefficient of 0.345. \citet{littlefair2007} show that typical uncertainties in $U_{w}$ are
$\sim5$\%, which leads to uncertainties in $R_{w}/a$ of $\sim1$\%. These errors have negligible
impact on our final system parameters.
As well as the parameters described above, the model also provides an estimate of the flux contribution
from the white dwarf, bright spot, accretion disc and donor. The white dwarf temperature and distance are
found by fitting the white dwarf fluxes from our model to the predictions of white dwarf model atmospheres
\citep{bergeron1995}, as shown in Fig. \ref{fig:wd_colours}. We find that with the exception of
CTCV 2354, all of the systems analysed lie near, or within, the range of white dwarf colours allowed
by the atmosphere models of \citet{bergeron1995}, although the systems do not always lie near the track
for the appropriate mass and radius of the white dwarf. \citet{littlefair2008} compare the temperatures
derived using light curve fits to those found using SDSS spectra and GALEX (Galaxy Evolution Explorer)
fluxes for a small number of systems and conclude their white dwarf temperatures are accurate to $\sim$1000K.
The systems examined by Littlefair et al. (2008) are all found to lie close to the Bergeron tracks; it
is likely that systems that lie far from the tracks are less accurate. We note our temperatures
have larger uncertainties than those of \citet{littlefair2008}. This is because our temperatures
take into account the uncertainty in white dwarf mass when comparing the white dwarf fluxes to the
models of \citet{bergeron1995}.
It is possible that our white dwarf colours are affected by contamination from the disc or bright spot,
or an unmodelled light source such as a boundary layer. If our white dwarf colours are incorrect, then
our derived white dwarf temperatures will be affected. Changing the white dwarf temperature will alter
$U_{w}$. Our model fitting measures $R_{w}/a$ and uses a mass-radius relationship to infer $M_{w}$, which
is then used to find the mass of the donor star. However, $U_{w}$ and $R_{w}$ are partially degenerate,
so $U_{w}$ therefore affects $R_{w}$ and $M_{w}$. $M_{w}$ is also affected by temperature changes because
the white dwarf mass-radius relationship is temperature dependent. The white dwarf temperature also affects
the luminosity of the system, and hence distance estimate. It is therefore important to quantify the effect
that incorrect white dwarf temperatures may have on distance estimates and our final derived system
parameters. To do this, we altered the white dwarf temperature by 2000K and performed the fitting procedure
described above on our best quality, white-dwarf dominated systems. For lower quality data, the random
errors dominate over any systematic errors, and thus changes to the best quality data represent a worst
case scenario. We find that changing the white dwarf temperature by 2000K changes $R_{w}/a$ by
less than 1$\sigma$. The white dwarf distance estimates change by 10-20pc. We therefore conclude any error
in white dwarf temperature that may occur does not affect our final system parameters by a significant amount.
We note here that our moedelling does not include treatment of any boundary layer around the white dwarf,
and assumes all of the white dwarf's surface is visible. Either effect could lead to systematic uncertainty
in our white dwarf radii \citep{wood1986}.
A Markov Chain Monte Carlo (MCMC) analysis was used to adjust all parameters bar $U_W$.
MCMC analysis is an ideal tool as not only does it provide a robust method for quantifying
the uncertainties in the various system parameters, it is more likely to converge on the global minimum
$\chi^{2}$ rather than a local minimum $\chi^{2}$. We refer the reader to \citet[][]{ford2006},
\citet[][]{gregory2007} and references therein for excellent overviews of MCMC chains and Bayesian
statistics and limit ourselves to a simple overview.
MCMC is a random walk process where at each step in the chain we draw a set of model parameters from a normal,
multi-variate distribution. This is governed by a covariance array, which we estimate from the initial stages
of the MCMC chain. The step is either accepted or rejected based on a transition probability, which is a function
of the change in $\chi^{2}$. We adopt a transition probability given by the Metropolis-Hastings (M-H) rule, that
is $P = \exp^{-\Delta\chi^{2}/2}$. The sizes of the steps in the MCMC chain are multiplied by a scale factor, tuned
to keep the acceptance rate near 0.23, which is found to be the optimal value for multi-variate chains such as
these \citep{roberts1997}.
A typical MCMC chain included some 700,000 steps, split into two, 350,000 step sections. The first section is used to
converge {\it towards} the global minimum and estimate the covariance matrix (known as the burn-in phase).
The second section fine tunes the solution by sampling areas of parameter space around the minimum. In doing so,
we also produce a robust estimation of our uncertainties. Together, these steps are usually sufficient to enable the
model to converge on the statistical best fit, regardless of the initial starting parameters.
While implementing the MCMC code, we discovered a bug in our original code. The re-binning code used to average
several light curves together mistreated the widths of the bins, which in turn affected the trapezoidal integration
of the model over these bins. The direct result was that in cases of heavy binning, such as systems with heavy flickering
or where several light curves had been averaged together (e.g. SDSS 1502), the white dwarf radius, $R_{w}/a$, was underestimated.
The exact amount depended on the level of binning used. This consequently resulted in an overestimate of the white dwarf
mass. Since the mass of the donor star, $M_{r}$, is related to the white dwarf mass $M_{w}$ by $M_{r} = qM_{w}$, we were
also left with an overestimate of the donor mass. This problem affects all of our previously published eclipsing-CV papers
(Feline et al. 2004a, 2004b; Littlefair et al. 2006a, 2006b, 2007, 2008) \nocite{feline2004a, feline2004b, littlefair2006a,
littlefair2006b, littlefair2007, littlefair2008} by differing amounts. However, in most cases re-modelling provides new system
parameters that are within $1-2 \sigma$ of our original results, with only two exceptions (see section \ref{sec:notes}).
The new results are presented in Table \ref{table:system_params}.
For each system we ran an MCMC simulation on each phase-folded $u'$, $g'$, $r'$ or $i'$ light curve from an arbitrary
starting position. Exceptions include CTCV 1300, where each night of observations was fit individually, and SDSS 1152
and SDSS 1501, for which we only calculated fits in the $g'$ and $r'$ bands due to $u'$-band data of insufficient quality
to constrain the model. Where no $u'$ band MCMC fit could be obtained, we fit and scaled the $g'$ band model to the $u'$
band light curves without $\chi^{2}$ optimisation. This allows us to estimate the white dwarf flux in the $u'$ band, and
thus estimate the white dwarf temperature. In the case of SDSS 1501, we also fit a different data set to the 2006 WHT data
of \citet{littlefair2008}. We fit
our model to the single light curve dated 2004 May 17. This 2004 data was not fit by Littlefair et al. (2008) as the simplex
methods used gave a seemingly good fit to the 2006 data. Despite appearing to have converged to a good fit, the MCMC
analysis revealed that the 2006 data does not constrain the model, most likely due to the very weak bright spot features.
The 2004 data shows much clearer and well-defined bright-spot features than the 2006 data (see Fig.1 of Littlefair et al.
2008), and so despite only having one eclipse (and thus lower signal-to-noise) it is favoured for the fitting process.
In general our fits to each system are in excellent agreement with the light curves (see Fig. \ref{fig:eclipses}),
giving us confidence that our new models accurately describe each system.
To obtain final system parameters we combine our MCMC chains with Kepler's $3^{rd}$ law, the orbital period, our
derived white dwarf temperature, and a series of white dwarf mass-radius relationships. We favour the relationships
of \citet{wood1995}, because they have thicker hydrogen layers which may be more appropriate for CVs.
However, they do not reach high enough masses for some of our systems. Above $M_{w} = 1.0 M_{\odot}$,
we adopt the mass-radius relationships of \citet{panei2000}. In turn, these models do not extend beyond
$M_{w} = 1.2 M_{\odot}$; above this mass we use the \citet{hamada1961} relationship. No attempt is made to
remove discontinuities from the resulting mass-radius relationship.
We calculate the mass ratio $q$, white dwarf mass $M_{w}/M_{\odot}$, white dwarf radius $R_{w}/R_{\odot}$, donor
mass $M_{r}/M_{\odot}$, donor radius $R_{r}/R_{\odot}$, inclination $i$, binary separation $a/R_{\odot}$ and radial velocities
of the white dwarf and donor star ($K_{w}$ and $K_{r}$, respectively) for each step of the MCMC chain. Since each step of the
MCMC has already been accepted or rejected based upon the Metropolis-Hastings rule, the distribution function for each parameter
gives an estimate of the probability density function (PDF) of that parameter, given the constraints of our eclipse data.
We can then combine the PDFs obtained in each band fit into the total PDF for each system, as shown in Fig. \ref{fig:pdfs}.
We note that most systems have system parameters with a Gaussian distribution with very little asymmetry. Our adopted value
for a given parameter is taken from the peak of the PDF. Upper and lower error bounds are derived from the 67\% confidence
levels. For simplicity, since the distributions are mostly symmetrical, we take an average of the upper and lower error bounds.
The final adopted system parameters are shown in Table \ref{table:system_params}, although Fig. \ref{figure:wd_masses} and
\ref{figure:models} show the true 67\% confidence levels for the white dwarf mass and donor mass respectively, for each system.
\begin{figure*}
\centering
\includegraphics[scale=0.80, trim=0 0 0 0]{fig2_lc1.eps}
\caption{The phased-folded $u'g'r'$ or $u'g'i'$ light curves of the CVs listed in Table \ref{table:system_params},
fitted using the model outlined in section \ref{sec:pme}.
The data (black) are shown with the fit (red) overlaid and the residuals plotted below (black). Below are the
separate light curves of the white dwarf (dark blue), bright spot (light blue), accretion disc (green) and the
secondary star (purple). Data points omitted from the fit are shown in red. $\chi^{2}$ values for each fit, together
with the number of degrees of freedom (DOF) are also shown.}
\label{fig:eclipses}
\end{figure*}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.80, trim=0 0 0 0]{fig2_lc2.eps}
\end{center}
\contcaption{}
\end{figure*}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.80, trim=0 0 0 0]{fig2_lc3.eps}
\end{center}
\contcaption{}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[scale=0.80, trim=0 0 0 0]{fig3_colours.eps}
\caption{The white dwarf colours derived from our model fitted together with the white dwarf models of \citet{bergeron1995}.
From top to bottom, each curve represents log g = 9.0, 8.5, 8.0, 7.5 and 7.0 respectively. The measured white dwarf colours
are shown here in red, and are used to derive the white dwarf temperature, which in turn is used to correct the white dwarf
mass-radius relationships used later to obtain the final system parameters.}
\label{fig:wd_colours}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[scale=0.80]{fig4_pdfs1.eps}
\caption{The normalised probability density functions for each system, derived using the MCMC chains, orbital period
and the mass--radius relationships of \citet{wood1995}, \citet{panei2000} and \citet{hamada1961}, at the appropriate
white dwarf temperature.
The red curve represents the $r'$ or $i'$ band fit, the green represents the $g'$ band, and blue curve (where present) represents
the $u'$ band. The black represents the total, combined PDF. Shown are the PDFs for mass ratio $q$, white dwarf radius
$R_{w}/R_{\odot}$, donor mass $M_{r}/M_{\odot}$, and inclination $i^{\circ}$.}
\label{fig:pdfs}
\end{figure*}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.80]{fig4_pdfs2.eps}
\end{center}
\contcaption{}
\end{figure*}
\begin{table*}
\vbox to220mm{\vfil Table 3 (system parameters) goes here (landscape). See separate file (table3params).
\caption{}
\label{table:system_params}
\vfil}
\end{table*}
\subsection{Notes on individual systems}
\label{sec:notes}
\subsubsection{CTCV 1300}
\label{sec:ctcv1300_2010}
As noted in section \ref{sec:pme}, the two eclipses of CTCV 1300 were modelled individually due to the heavy pre-eclipse
flickering present in each light curve. The $u'g'r'$ eclipses from the night of 2007 June 10 gave consistent results, as
did the $u'g'i'$ eclipses from the night of 2007 June 13. However, results from the individual nights were not consistent
with each other, and we are presented with two distinct solutions which are the result of heavy pre-eclipse flickering
altering the shape of the bright spot ingress feature. This in turn gives two very different values for the mass ratio,
$q$. Encouragingly we find our white dwarf masses and radii are consistent between nights. To derive the final system
parameters we use the PDFs as outlined in section \ref{sec:pme} and then take an average of the solution from each
night. The error is taken as the standard deviation between the two values.
In Fig. \ref{fig:ctcv1300_2010} we see a $g'$-band eclipse from June 2010 together with a modified version of the
model obtained from the 2007 dataset. This new model is found by starting from an average of the two 2007 fits and
using a downhill simplex method to vary all parameters bar $q$, $\Delta\phi$, $R_{w}/a$ and $U_{w}$. These parameters
should not change with time, and so the simplex fit will confirm if our bright spot positions and white dwarf dwarf
radius are correct, and thus if our system parameters are reliable. We use a simplex method for two reasons; firstly
since we are performing a consistency check and are not extracting system parameters from the fitting process we do
not require a full MCMC analysis. Secondly, the bright spot flux appears to have reduced significantly, and the
strength of the bright spot ingress feature means that we cannot constrain a full fit using the MCMC model used previously.
The model confirms the bright spot flux has decreased considerably, although an orbital hump is still visible. The white
dwarf flux remains almost unchanged, although the disc appears brighter. The fit to the data is good, indicating that
the models derived from the 2007 data (and used to derive our system parameters) are reliable.
\begin{figure}
\centering
\includegraphics[scale=1.10,trim=0 0 0 0]{fig5_ctcv1300.eps}
\caption{Our June 2010 $g'$ band eclipse of CTCV 1300, together with a modified model found using the 2007 data.
Starting from an average of the two 2007 models, we ran a downhill simplex fit varying all parameters bar $q$,
$\Delta\phi$, $R_{w}/a$ and $U_{w}$.}
\label{fig:ctcv1300_2010}
\end{figure}
\subsubsection{CTCV 2354}
\label{sec:ctcv2354_2010}
In section \ref{sec:2354} we noted that the shape of the light curve in Fig. \ref{figure:lightcurves} indicated
possible bright spot egress features around phases 1.060 and 1.080. Fig. \ref{fig:eclipses} shows that our model has
fit the bright spot egress feature at phase 1.080. Given the strength and shape of the bright spot features and general
scatter present in the light curve we cannot be certain if the bright spot positions have been correctly identified by our
model, and thus there is some element of doubt as to the value obtained for our mass ratio and thus donor mass.
It is at this point we draw the readers attention to our 2010 data, shown in Fig. \ref{figure:ctcv2354_2009_2010}.
The eclipse dated 2010 May 3 (centre panel) shows clear bright spot ingress and egress features, with a clear
orbital hump visible from phases 0.70--0.95. The system is much brighter in this state than the 2007 data previously
modelled, in part due to a dramatic increase in bright spot flux. As with CTCV 1300 we carried out a downhill
simplex method to vary all parameters bar $q$, $\Delta\phi$, $R_{w}/a$ and $U_{w}$. Our fit to the eclipse of May 3
is especially pleasing, as its excellent agreement with the light curve confirms that our 2007 model correctly identified
the bright spot egress feature and thus the mass ratio obtained is reliable.
Fig. \ref{figure:ctcv2354_2009_2010} also shows a single eclipse observed just one month later (June 2010, right panel),
and six eclipses averaged together from June 2009 (left panel). Both of these datasets are fit with a downhill simplex model
as above. The June 2009 and June 2010 datasets are in stark contrast to the May 2010 data, with the bright spot features
appearing extremely faint (2009), or seemingly non-existent (June 2010). The disc flux in the June 2010 data appears to
have increased significantly, giving rise to a distinct ``u'' shape. The rapid change in bright spot and disc light curves
over such short (1 month - 1 year) time scales suggests that the disc is highly unstable.
\begin{figure*}
\centering
\includegraphics[scale=0.8,trim=0 0 0 0]{fig6_ctcv2354.eps}
\caption{Our $g'$ band observations of CTCV 2354 from 2009 (left), 2010 May 3 (centre) and 2010 June 6 (right),
fit with a modified version of the model obtained from the 2007 data as described in section \ref{sec:ctcv2354_2010}.}
\label{figure:ctcv2354_2009_2010}
\end{figure*}
\subsubsection{DV UMa}
Our donor mass derived for DV UMa has increased by $6.1\sigma$ ($\Delta$M=0.039$M_{\odot}$) from that published by Feline et
al. (2004b). \nocite{feline2004b} Close inspection of the original fit reveals that the bright spot features are fit
poorly by the old bright spot model. This arises because the old bright spot model could not describe the complex bright
spot profile present and an innacurate value of the mass ratio is found as a result. Our new bright spot model is much
better in this respect, and is able to take into account a wider variety of geometric effects and orientations. Given
that our white dwarf radius is consistent with that of Feline et al. (2004b), this seems the most likely cause of such
a large change. It is worth noting that our new donor masses for both DV UMa and XZ Eri, are both consistent with the
masses obtained by Feline et al. (2004b) using the derivative method, which, unlike our parameterised model, does not
make any attempt to recreate the bright spot eclipse profile (e.g. Wood et al. 1986; Horne et al. 1994; Feline et al.
2004a; Feline et al. 2004b). \nocite{wood1986, horne1994, feline2004a, feline2004b}
\subsubsection{SDSS 1502}
Our new fits to SDSS 1502 decrease the donor mass by $2.9\sigma$ ($\Delta$$M_{r}=0.012M_{\odot}$) from that of
\citet{littlefair2008}. Our mass
ratio and inclination are consistent with those of \citet{littlefair2008}, however
our white dwarf radius, $R_{w}$, has increased by $13$ percent ($3.4\sigma$). We believe the primary reason for
this change was that the original fit was heavily binned, and thus more susceptible to the bug outlined in
section \ref{sec:pme}.
\subsubsection{SDSS 1501}
\label{sec:SDSS1501}
The most important change of all of our re-modelled systems is for that of SDSS 1501. Whilst our donor mass
has only increased by $1.9\sigma$ from that of \citet{littlefair2008}, we note that our uncertainties are large
($\sigma$$M_{r}=0.010M_{\odot}$) and the mass difference is large enough to take this system from being a
post-period-bounce system, to a pre-period-bounce system.
Although our errors do not formally rule out the possibility of SDSS 1501 being post-period-bounce, the donor's position with
respect to the evolutionary tracks shown in Fig. \ref{figure:models} strongly favours that of a pre-period-bounce
system. Such a change arises from a difference in bright spot positions between our model and that of \citet{littlefair2008},
which in turn affect the mass ratio obtained. The data used by \citet{littlefair2008} shows a very weak bright spot
ingress feature. With the improvements made to the modelling process resulting from the introduction of MCMC, it is
clear that the 2006 data used by \citet{littlefair2008} does not constrain the mass ratio, $q$, tightly enough.
In contrast, the 2004 data shows much clearer bright spot features, and we therefore favour this over the 2006 data
as discussed in section \ref{sec:pme}.
\section{Discussion}
\label{sec:discussion}
\subsection{White dwarf masses}
Population studies by \citet[][]{willems2005} predict that between 40 and 80 percent of CVs are born with He-core white
dwarfs ($M_{w} \lesssim 0.50 M_{\odot}$) and therefore He-core white dwarfs (He-WDs) are expected to be common amongst CV
primaries. It is surprising then that out of our sample of 14 systems, we observe no He-WDs. Of all of our objects, SDSS
1152 is found to have to the lowest white dwarf mass with $M_{w} = 0.560\pm0.028$. The mass distribution of \citet{kepler2007}
for SDSS white dwarfs suggests He-WDs have a typical mass of $\sim$0.38$M_{\odot}$. The most massive He-WDs are thought to
form from single RGB stars, which due to extreme mass loss are able to avoid the Helium flash; \citet{dcruz1996} consider
models with a range of mass loss rates on the RGB and manage to produce He-WDs with masses up to $\sim$0.48$M_{\odot}$. It
is likely that this represents an upper limit to the mass of He-WDs and hence SDSS 1152 is too massive to be a viable
candidate for a He-core white dwarf.
We find that our white dwarf masses are not only too massive to be He-WDs, but are also well in excess of the average
mass for single DA white dwarfs. Using the same method as \citet{knigge2006}, we calculated the average white dwarf mass
of our entire sample to be $M_{w} = 0.81\pm0.04 M_{\odot}$, with an intrinsic scatter of $0.13 M_{\odot}$.
In comparison, \citet{liebert2005} find the mean mass of DA white dwarfs to be $M_{w}\sim$0.603 $M_{\odot}$,
while \citet{kepler2007} find a mean mass of $M_{w} = 0.593\pm0.016 M_{\odot}$.
Our study thus supports previous findings \citep[e.g.][]{warner1973, warner1976, ritter1976, ritter1985, robinson1976,
smith1998, knigge2006} that white dwarfs in CVs are on average
much higher in mass than single field stars. Like \citet{littlefair2008}, we compare our masses to the average mass of
$M_{w} = 0.73\pm0.05 M_{\odot}$ for white dwarfs in CVs below the period gap \citep{knigge2006} and find that our white
dwarf masses are generally much higher. This is especially so for systems $P_{orb} \leq 95$ mins, where we find a mean
white dwarf mass of $M_{w} = 0.83\pm0.02 M_{\odot}$, with an intrinsic scatter of $0.07 M_{\odot}$. Some our white dwarf
masses are revised, and have accordingly moved downwards in mass compared to \citet{littlefair2008}, but as Fig.
\ref{figure:wd_masses} shows, for $P_{orb} \leq 95$ mins, 8 out of 9 systems are more massive than the average found by
\citet{knigge2006}, with SDSS 1502 the only exception. On the same plot, we also plot the dispersion of masses as found
by \citet{knigge2006} and the mean mass of single SDSS white dwarfs found by \citet{kepler2007}. Note that the
\citet{knigge2006} sample contains three systems also included in our study: OU Vir, DV UMa and XZ Eri, using the
old mass determinations of Feline et al. (2004a, 2004b). We see most of our masses are within the dispersion found,
indicating that no individual white dwarf mass is unusual. However 8 out of 9 systems above average does seem anomalously
high, considering that if we model the white dwarf masses as a Gaussian distribution, the probability of such an occurrence
is less than 2 percent (independent of the actual mean or variance). Such difference between our sample and that of
\citet{knigge2006} is concerning, and it is therefore desirable to consider the selection effects, considering the
majority of our short period systems are all SDSS objects \citep{szkody2004, szkody2005, szkody2006, szkody2007}.
The majority of SDSS CVs found are generally rejected quasar candidates with a limiting magnitude of $g'=19-20$,
and are initially selected for follow up on the basis of $u'$-$g'$ colour cuts (see \citet{gaensicke2009}
for a more in depth description). \citet{littlefair2008} show that systems with $M_{w} \geq0.50 M_{\odot}$ are blue enough
to pass the SDSS colour cuts and conclude that selection effects such as these are unlikely to explain the high
mass bias of our white dwarf sample. However, the majority of the systems we have studied are close ($g'$$\sim17.5-19.5$)
to the $g'=19$ limit of the SDSS survey. This raises the possibility that the SDSS sample only finds the brightest of the
short period CVs. \citet{ritter1986} have shown that CVs with high mass white dwarfs are brighter than their low mass
counterparts. This suggests there maybe some bias towards high mass dwarfs. However, this conclusion is not appropriate
to the systems studied here: \citet{ritter1986} only consider the effects of accretion luminosity whereas in most of
our systems the white dwarf considerably outshines the accretion disc.
\citet{zorotovic2011} consider selection effects in white dwarf dominated SDSS systems with a variety of different
mass transfer rates and conclude there is actually bias {\it against} high mass white dwarfs. They find a 0.90$M_{\odot}$
white dwarf is approximately 0.15 magnitudes fainter than a 0.75$M_{\odot}$ white dwarf, which corresponds to a decreased
detection efficiency of $\sim20\%$. This suggests that our finding of high mass white dwarfs at short orbital periods is
not due to selection effects, and is in fact a true representation of the intrinsic mass distribution of CVs. However,
this analysis only considers white dwarf luminosity; in some of our systems the bright spot features are
prominent, and contribute significantly to the overall flux of the system. It remains possible that the finding of
very high white dwarf masses for our short period CVs is due to selection effects. However, a full and thorough quantification
of any bias would require detailed calculations of the luminosity of white-dwarf dominated systems (including bright
spot emission) plus an investigation of the selection effects in the SDSS sample. Such an analysis is beyond the
scope of this paper.
Our results have important consequences for the modelling of nova outbursts and their impact on the long-term evolution
on CVs. Typical calculations show that the mass of the white dwarf decreases by between 1 and 5 percent per 1000 nova cycles
\citep[e.g.][]{yaron2005, epelstain2007}. The dominance of high-mass white dwarfs in our sample of short period systems
suggests that any white dwarf erosion due to nova explosions must be minimal, or that not all of the accreted matter is
ejected during nova ignition, resulting in the white dwarf mass {\it increasing} over time. This could, in principle
enable the white dwarfs in cataclysmic variables to grow in mass until they reach the Chandrasekhar limit.
\begin{figure*}
\centering
\includegraphics[scale=0.6, angle=-90]{fig7_wds.eps}
\caption{White dwarf mass as a function of orbital period. The mean white dwarf mass for systems below the period gap,
as found by \citet{knigge2006} is shown with a solid line, along with the associated intrinsic scatter (dashed line).
The mean white dwarf mass in single stars as found by \citet{kepler2007} is shown by a dotted line.}
\label{figure:wd_masses}
\end{figure*}
\subsection{Period bounce}
Population synthesis models for cataclysmic variables \citep[e.g.][]{kolb1993, willems2005} all predict that large
numbers of the CV population ($\sim$15 - 70 percent) have evolved past the period minimum. This has always been in
stark contrast to observations, possibly in part due to selection effects \citep[e.g.][]{littlefair2003}.
\citet{littlefair2006b, littlefair2007, littlefair2008} identified four systems (SDSS 1035, SDSS 1507, SDSS 1433,
SDSS 1501) with donors below the sub-stellar limit, three of which are likely to be post-period-bounce CVs (SDSS
1035, SDSS 1501 and SDSS 1433). Our subsequent re-analysis gives three systems (SDSS 1035, SDSS 1507, SDSS 1433) with
donors below the sub-stellar limit, two of which (for reasons outlined below) we believe are post-period-bounce
CVs (SDSS 1035 and SDSS 1433). SDSS 1501, which no longer features as a post-period-bounce system, is discussed in
greater detail in section \ref{sec:SDSS1501}.
\citet{sirotkin2010} claim that SDSS 1433 cannot be considered a post-period-bounce object since the mass transfer
rates and donor star temperatures implied are too high. The mass transfer rate is
found using an estimate of the white dwarf temperature \citep{townsley2009}, while the donor star temperature
is inferred using a semi-empirical relationship that is also dependent on the white dwarf temperature. The white
dwarf temperature used by \citet{sirotkin2010} is that derived by \citet{littlefair2008} from model fitting.
We believe, at least in this case, that using $\dot{M}$ and $T_{2}$ is an unreliable test of the evolutionary status
of CVs donors, since accurate determinations of the white dwarf temperature are difficult to obtain. Of all the
system parameters we have derived, the white dwarf temperatures are the least well constrained, and this does not take
into account systematic errors. Since the white dwarf temperature
is found using the flux from just three colours, and our model does not include all possible sources of luminosity
(e.g. a boundary layer), there is a good chance our white dwarf temperatures are affected by systematic errors at some level,
as discussed in section \ref{sec:pme}. Instead, we focus on the donor star mass, $M_{r}$.
If the angular momentum loss rate is similar for systems with identical system parameters, we expect all CVs to
follow very similar evolutionary tracks
with a single locus in the mass-period relationship (and by analogy, mass-radius relationship) for CV donors, as
shown in Fig. \ref{figure:models}. The empirical donor star mass-radius relationship derived by \citet{knigge2006}
shows that a single evolutionary track does very well at describing the observed $M_{2}-P_{orb}$ relationship, although
the shape of that relationship is poorly constrained at low masses. A single evolutionary path also explains the
presence of the ``period spike'', a long sought after feature in the orbital period distribution recently identified
by \citet{gaensicke2009}. We therefore expect there to be a unique donor mass corresponding to the minimum orbital
period, below which an object becomes a period-bouncer. The exact mass at which this occurs is very uncertain, and
does not necessarily correspond to the sub-stellar limit \citep{patterson2009}. From the empirical work of \citet{knigge2006},
the best estimate for $M_{bounce}$ is $M_{r} = 0.063 \pm 0.009 M_{\odot}$. Three of our systems (SDSS 1035, SDSS 1433
and SDSS 1507) fall well below this value, although SDSS 1507 is an unusual system, and is discussed in the following
section. As in \citet{littlefair2008}, we do not include it in our sample of post-period minimum CVs. We therefore have two
strong candidates for post-period minimum CVs (SDSS 1035 and SDSS 1433) from our total sample of
14 CVs (nine of which are SDSS systems). From this, we estimate that $14 \pm 7$ percent of all CVs below the period
gap, and $22 \pm 11$ percent of all short period CVs ($P_{orb} \leq 95$ mins) have evolved
past the period minimum. These findings are consistent, albeit to a crude approximation given our small sample of objects,
with current population synthesis models. Since all of our short period systems are SDSS CVs, we cannot rule out selection
effects, but \citet{gaensicke2009} have shown that the number of period minimum CVs found within the SDSS is broadly
consistent with other surveys, allowing for normalisation of survey volumes.
\subsection{SDSS 1507}
The orbital period of SDSS 1507 is far below the well-defined period minimum and thus the nature of this system is of
great interest to theorists and observers. It is possible that this system represents the true orbital period minimum
as predicted by \citet{kolb1999}. However, if this is indeed the case, we would expect a large number of systems between
orbital periods of 67 minutes, and 83 minutes where the period spike is observed \citep{gaensicke2009}. These systems are
not observed, and hence it is likely that some other mechanism is responsible. \citet{littlefair2007} speculate that this
system was either formed directly from a white dwarf/brown dwarf binary, while \citet{patterson2008} argue that the system
could be a member of the halo. Both derive system parameters, and both obtain distance estimates to the system.
Our derived system parameters are consistent with those of \citet{littlefair2007} and \citet{patterson2008}, within
uncertainties. Our distance estimate is in excellent agreement with \citet{littlefair2007}, which is not suprising
since we both calculate the distance using the same methods and dataset. However, our new distance estimate still
places the system nearer than that of \citet{patterson2008}. \citet{patterson2008} obtain a lower limit to the
distance using parallax. The parallax value implies a distance $d > 175$ pc, which taken alone,
is consistent with our estimate of $d=168\pm12$ pc. Patterson combines his parallax with a range of other
observational constraints using Bayesian methods to
yield a final distance estimate of $d=230\pm40$ pc. If our distance of $d=168\pm12$ pc is nearer the true distance,
then combining with Patterson's proper motion measurement of $0.16 ''/yr$ yields a transverse velocity of $d=128\pm9$
kms$^{-1}$. This lower transverse velocity is still very much an outlier in the distribution of 354 CVs shown in Fig.1 of
\citet{patterson2008}. Therefore, regardless of which distance is correct, the proper motion of SDSS 1507 still
supports halo membership.
\subsection{Exploring the standard model of CV evolution}
Fig. \ref{figure:models} shows the evolutionary models of \citet{kolb1999} calculated with enhanced mass-transfer
rates. Also shown is a model with 50 percent star spot coverage on the surface of the donor. Positions of the period
minimum, and period gap as found by \citet{knigge2006} are also shown. Mass determinations for all systems presented
here are included. We see that the standard theoretical models are a poor fit to the data. For a given mass, the models of
\citet{kolb1999} significantly underestimate the orbital period, and thus the donor radii.
Models with enhanced mass transfer rates and star spot coverage do rather better at reproducing the observed donor
masses, although the general scatter of short period systems makes choosing between these difficult. This is in
line with the conclusions of \citet{littlefair2008}. The models begin to diverge significantly at orbital
periods greater than 100 minutes. Unfortunately, in this regime there are few systems with precisely known donor masses.
Clearly, we require more mass determinations for systems with orbital periods between 100 and 130 minutes.
\begin{figure*}
\centering
\includegraphics[scale=0.6, angle=0]{fig8_donors.eps}
\caption{The $M_{2}-P_{orb}$ relationship for our dataset. Mass determinations for all systems using {\sc ultracam}
data are included. Our three new systems, in addition to other objects of particular interest are labelled.
The evolutionary models of \citet{kolb1999} calculated with different mass-transfer rates are shown with red
(dashed) and blue (dot-dashed) lines. A model with 50 percent star spot coverage on the surface of the donor
is shown with an orange (dotted) line. The solid (black) line shows the empirical mass-radius relationship as
found by \citet{knigge2006}. The position of the period minimum and period gap, as found by \citet{knigge2006},
are also shown.}
\label{figure:models}
\end{figure*}
\section{Conclusions}
\label{sec:conclusion}
We present high-speed, three-colour photometry of a sample of 14 eclipsing CVs. Of these CVs, nine are short period
($P_{orb} \leq 95$ minutes), and one is within the period gap. For each of the 14 objects we determine the system
parameters by fitting a physical model of the binary to the observed light curve by $\chi^{2}$ minimisation. We find
that two of our nine short period systems appear to have evolved past the period minimum, and thus
supports various assertions that between 15 and 70 per cent of the CV population has evolved past the orbital period
minimum. The donor star masses and radii are not consistent with model predictions, with the majority of donor stars
being $\sim$10 per cent larger than predicted. Our derived masses and radii show that this can explained by either
enhancing themass transfer rate or modifying the stellar physics of the donor star to take into account star spot coverage.
Unfortunately, we still lack enough precise donor masses between orbital periods of 100 and 130 minutes to choose between
these alternatives.
Finally, we find that the white dwarfs in our sample show a strong tendency towards high masses. The high mass dominance
implies that the white dwarfs in CVs are not significantly eroded by nova outbursts, and may actually increase over several nova
cycles. We find no evidence for He-core white dwarfs within our sample, despite predictions that between 40 and 80 percent
of short period CVs should contain He-core white dwarfs.
\section{Acknowledgements}
We would like to thank our referee, Joe Patterson for his useful comments. We also
thank Christian Knigge for useful discussions on white dwarf bias and selection effects. CDJS acknowledges
the support of an STFC PhD. SPL acknowledges the support of an RCUK Fellowship. CMC and TRM are supported under grant
ST/F002599/1 from the Science and Technology Facilities Council (STFC). {\sc ultracam} and
{\sc ultraspec} are supported by STFC grant ST/G003092/1. This research has made use of NASA's Astrophysics Data System
Bibliographic Services. This article is based on observations made with {\sc ultracam} mounted on the Isaac Newton Group's
WHT, and {\sc ultracam} and {\sc ultraspec} mounted on the European Southern Observatory's NTT and VLT telescopes.
\bibliographystyle{mn2e}
|
\section{Introduction}
The notion of Lyapunov function plays a very important role in
design and verification of dynamical systems, in particular, in
performance analysis, stability analysis and controller synthesis of
complex dynamical and controlled systems
\cite{Haddad-Che,Khalil,Li-Slotine}. \,In recent years, people
realized that the notion is quite helpful to safety verification of
hybrid and cyber-physical systems as well \cite{Tabuada}.
However, the following two issues hinder the application of Lyapunov
functions in practice. Firstly, it is actually not necessary for
the first-order Lie derivative of a Lyapunov function to be strictly
negative to guarantee asymptotic stability, which is shown by
\emph{LaSalle's Invariance Principle} \cite{Khalil}. Such a
condition could limit to scale up the method. Secondly, in general
there is no effective way so far to find Lyapunov functions,
although many methods have been proposed by different experts using
their field expertise.
To address the above two issues, in this paper, we first generalize
the standard concept of Lyapunov function to \emph{relaxed Lyapunov
function} (RLF) for asymptotic stability analysis. Compared with the
conventional definition of Lyapunov function, the first non-zero
higher order Lie derivative of RLF is required to be negative,
rather than its first-order Lie derivative. Such a relaxation
extends the set of admissible functions that can be used to prove
asymptotic stability.
Another contribution of this paper is that we present a complete
method to automatically discovering polynomial RLFs for polynomial dynamical systems (PDSs).
The basic idea of our method is to predefine a parametric polynomial as a template of
RLF first, and then utilize
the Lie derivatives of the template at different orders to
generate constraint on the parameters, and
finally solve the resulting constraint.
Our method is complete in the sense that it
is able to generate all polynomial RLFs by enumerating all polynomial templates for any PDS.
\newline
\noindent {\bf{Related Work.}} In \cite{RS}, the same terminology
``relaxed Lyapunov function" is used, with a different definition.
The idea of applying higher order Lie derivatives to analyze
asymptotic stability is not new. For example, in \cite{Butz,Mei-Ni}
the authors resorted to certain linear combinations of higher order
Lie derivatives with non-negative coefficients such that the
combination is always negative. This method could be included in the
framework of vector Lyapunov functions method
\cite{Matrosov,Ner-Had}. Our method is essentially different from
theirs because an RLF only requires its first non-zero higher order
Lie derivative to be negative.
In the literature, there is a lot of work on constructing Lyapunov
functions. For instance, in
\cite{Gurvits,Liberzon,Agrachev-Liberzon}, methods for constructing
common quadratic Lyapunov functions for linear systems were
proposed, which were generalized in \cite{SNS} and
\cite{Vu-Liberzon} for nonlinear systems wherein the generated Lyapunov functions are not necessarily quadratic. Another useful technique is the linear matrix inequality (LMI) method introduced in \cite{Jo-Ran} and \cite{Pet-Ben}, which enables us to utilize the results of numerical optimization for discovering piecewise quadratic Lyapunov functions. Based on sums-of-squares (SOS)
decomposition and semi-definite programmng (SDP) \cite{Parrilo}, a
method for constructing piecewise high-degree polynomial and
piecewise non-polynomial Lyapunov functions was proposed in
\cite{Pra-Pap} and \cite{Pra-Pap1}. The SOS and SDP based method
was also used in \cite{TP} to search for control Lyapunov functions
for polynomial systems. In \cite{SXX}, the authors proposed a new
method for computing Lypunov functions for polynomial systems by
solving semi-algebraic constraint using their tool DISCOVERER
\cite{Xia07}. Approaches to constructing Lyapunov functions beyond
polynomials using radial basis functions were proposed in
\cite{Giesl2,Giesl3}.
Our method has the following features compared to the related work.
Firstly, it generates relaxed Lyapunov functions rather than
conventional Lyapunov functions. Secondly, it is able to discover
all polynomial RLFs by enumerating all polynomial templates
for any PDS, whereas the Krasovskii's method \cite{Kra} and Zubov's
method \cite{Zubov} can only produce Lyapunov functions of special
forms. Thirdly, the LMI method and SOS method are numerical, while
our method is symbolic, which means it could provide a
mathematically rigorous framework for the stability analysis of
polynomial dynamical systems.
\noindent {\bf{Structure:}} The rest of this paper is organized as
follows. In Section \ref{sec:foundations}, the
theoretical foundations are presented.
Section \ref{sec:RLF} shows a new criterion for asymptotic stability using the notion of relaxed Lyapunov functions.
In Section \ref{sec:main-result} we present a sound and complete
method and a corresponding algorithm for automatically discovering
polynomial RLFs on polynomial dynamical systems. The method is
illustrated by an example in Section \ref{sec:example}. Finally, we
conclude this paper and discuss possible future work in Section
\ref{sec:conclusion}.
\section{Theoretical Foundations}\label{sec:foundations}
In this section, we present the fundamental materials based on which
we develop our method.
\subsection{Polynomial Ideal Theory}
Let $\mathbb{K}$ be an algebraic field, and $\mathbb{K}[x_1, x_2,
\dots, x_n]$ denote the polynomial ring over $\mathbb{K}$.
Customarily, let $\xx$ denote the $n$-tuple $(x_1, \cdots, x_n)$.
Then $\mathbb{K}[x_1, x_2, \dots, x_n]$ can be written as
$\mathbb{K}[\xx]$ for short, and a polynomial in $\mathbb{K}[x_1,
x_2, \dots, x_n]$ can simply be written as $p(\xx)$ or $p$.
Particularly, $\mathbb{K}$ will be taken as the real field $\mathbb
R$ in this paper, and $\xx$ takes value from the $n$-dimensional
Euclidean space $\mathbb{R}^n$.
In our method we will use polynomials with undetermined
coefficients, called parametric polynomials or \emph{templates}.
Such polynomials are denoted by $p(\uu,\xx)$, where
$\uu=(u_1,u_2,\ldots,u_t)$ is a $t$-tuple of parameters. A
parametric polynomial $p(\uu,\xx)$ in $\mathbb{R}[x_1, x_2, \dots,
x_n]$ with real parameters can be seen equivalently as a regular
polynomial in $\mathbb{R}[u_1, u_2, \ldots, u_t, x_1, x_2, \ldots,
x_n]$. Given $\uu_0\in \mathbb{R}^t$, we call the polynomial
$p_{\uu_0}(\xx)$ resulted by substituting $\uu_0$ for $\uu$ in
$p(\uu,\xx)$ an \emph{instantiation }of $p(\uu,\xx)$.
The following are some fundamental results relative to
polynomial ideals, which can be found in \cite{clo}.
\begin{definition}
A subset $I\subseteq \mathbb{K}[\xx]$ is called an ideal iff
\begin{enumerate}
\item[(a)] $0\in I$;
\item[(b)] If $p(\xx), g(\xx)\in I$, then $p(\xx)+g(\xx)\in I$;
\item[(c)] If $p(\xx)\in I$, then $p(\xx)h(\xx)\in I$ for any $h(\xx)\in
\mathbb{K}[\xx]$.
\end{enumerate}
\end{definition}
It is easy to check that if $p_1, \cdots, p_m\in \mathbb{K}[\xx]$, then
\[\<p_1, \cdots, p_m\rg=\{\sum_{i=1}^{m}p_ih_i\mid \forall i\in [1,m].\, h_i \in
\mathbb{K}[\xx]\}\] is an ideal. In general, we say
an ideal $I$ is {\itshape generated} by polynomials $g_{1}, g_2,
\dots, g_k\in \mathbb{K}[\xx]$ if $I=\<g_{1}, g_2,
\dots, g_k\rg,$ where all $g_i$ for $i \in [1,k]$ are called {\itshape
generators} of $I$. In fact, we have
\begin{theorem}[Hilbert Basis Theorem]\label{basis}\ \
Every ideal $I$ $\subseteq \mathbb{K}[\xx]$ has a
finite generating set. That is, $I=\<g_{1}, g_2, \dots, g_k\rg$ for
some $g_{1}, g_2, \dots, g_k \in \mathbb{K}[\xx]$.
\end{theorem}
From this result, it is easy to see that
\begin{theorem}[Ascending Chain Condition]\label{ACC}\ \
For any ascending chain
\[I_1\subseteq I_2 \subseteq \cdots \subseteq I_m \subseteq \cdots\] of
ideals in polynomial ring $\mathbb{K}[\xx]$, there
must be an $N$ such that for all $m\geq N$, $I_m=I_N$.
\end{theorem}
\subsection{Dynamical Systems and Stability}
We summarize some fundamental theories of dynamical systems here.
For details please refer to \cite{Haddad-Che,Khalil,Li-Slotine}.
\subsubsection{Dynamical Systems}
We consider autonomous dynamical systems modeled by first-order ordinary differential equations
\begin{equation}\label{eq:ode}
\dot \xx= \fb(\xx) \enspace,
\end{equation}
where $\xx\in\mathbb{R}^n$ and $\fb$ is a vector function from $\mathbb R^n$ to $\mathbb R^n$, which is also called a vector field in $\mathbb R^n$.
In this paper, we focus on special nonlinear dynamical systems whose vector fields are defined by polynomials.
\begin{definition}[Polynomial Dynamical System]\label{dfn:ads}\ \
Suppose $\fb=(f_1,f_2,\cdots,f_n)$ in (\ref{eq:ode}). Then (\ref{eq:ode}) is called a \emph{polynomial dynamical system} (PDS for short) if for every $1\leq i\leq n$, $f_i$ is a polynomial in $\mathbb R[\xx]$.
\end{definition}
If $\fb$ satisfies the local Lipschitz condition, then given
$\xx_0\in \mathbb R^n$, there exists a unique solution $\xx(t)$ of
(\ref{eq:ode}) defined on $(a, b)$ with $a< 0< b$ s.t.
$$\forall t\in (a,b).\,{\ud \xx(t)\over \ud t} = \fb (\xx(t))\quad \mathrm{and}\quad \xx(0)=\xx_0\enspace .$$
We call $\xx(t)$ on $[0,b)$ the \emph{trajectory} of (\ref{eq:ode}) starting from initial point $\xx_0$.
Let $\sigma(\xx)$ be a function from $\mathbb R^n$ to $\mathbb R$. Suppose both $\sigma$ and $\fb$ are differentiable in $\xx$ at any
order $n\in \mathbb N$. Then we can inductively define the {\itshape
Lie derivatives} of $\sigma$ along $\fb$, i.e.
$L^k_{\fb}\sigma: \mathbb R^n\rightarrow \mathbb R$ for $k\in \mathbb N$, as follows:
\begin{itemize}
\item $L^0_{\fb} \sigma(\xx)=\sigma(\xx)$,
\item $L^k_{\fb} \sigma(\xx)=\left( \frac{\partial}{\partial \xx} L^{k-1}_{\fb} \sigma(\xx), \fb(\xx)\right)$, for $k>0$,
\end{itemize}
where $(\cdot, \cdot)$ is the inner product of two vectors, i.e. $(
\,(a_1, \ldots, a_n), ( b_1, \ldots, b_n )\,) =\sum_{i=1}^n a_ib_i.$
Polynomial functions are sufficiently smooth, so given a PDS
$\mathcal{P}$ and a polynomial $p$, the vector field $\fb$ of
$\mathcal{P}$ satisfies the local Lipschitz condition, and the
higher order Lie derivatives of $p$ along $\fb$ are well defined and
are all polynomials. For a parameterized polynomial $p(\uu,\xx)$,
we can define $L_{\fb}^k p(\uu,\xx): \mathbb R^n\rightarrow \mathbb
R$ by seeing $\uu$ as undetermined constants rather than variables.
In the sequel we will implicitly employ these facts.
\begin{example}\label{eg:Lie-derv}
Suppose $\fb=(-x,y)$ and $p(x,y)=x+y^2$. Then $L^0_{\fb} p=x+y^2$, $L^1_{\fb} p=-x+2y^2$, $L^2_{\fb} p=x+4y^2$, $L^3_{\fb} p=-x+8y^2$ .
\end{example}
\subsubsection{Stability}
The following are classic results of stability of dynamical systems in the sense of Lyapunov.
\begin{definition}\
A point $\xx_e\in \mathbb R^n$ is called an {\itshape equilibrium or
critical point} of (\ref{eq:ode}) if $\fb(\xx_e)=\mathbf{0}$.
\end{definition}
We assume $\xx_e=\mathbf{0}$ w.l.o.g from now on.
\begin{definition}\
Suppose $\mathbf{0}$ is an equilibrium of (\ref{eq:ode}). Then
\begin{itemize}
\item
$\mathbf{0}$ is called Lyapunov stable if for any
$\epsilon>0$, there exists a $\delta>0$ such that if $\|
\xx_0\|<\delta$, then for the corresponding solution $\xx(t)$ of (\ref{eq:ode}), $\|\xx(t)\|<\epsilon$ for all $t\geq 0$.
\item
$\mathbf{0}$ is called asymptotically stable if it is
Lyapunov stable and there exists a $\delta>0$ such that for any $\|
\xx_0\|<\delta$, the
corresponding solution $\xx(t)$ of (\ref{eq:ode}) can be extended to infinity and $\lim_{t\rightarrow
\infty}\xx(t)=\mathbf{0}$.
\end{itemize}
\end{definition}
Lyapunov first provided a sufficient condition, using so-called \emph{Lyapunov
function}, for the Lyapunov stability as follows.
\begin{theorem}[Lyapunov Stability Theorem]\label{Lya}\
Suppose $ \mathbf{0}$ is an equilibrium point of
(\ref{eq:ode}). If there is an open set $ U\subset \mathbb R^n$ with
$\mathbf{0} \in U$ and a continuous differentiable function $V: U\rightarrow \mathbb R$
such that
\begin{enumerate}
\item[(a)] $V(\mathbf{0})=0$,
\item[(b)]$V(\xx)>0$ for all $\xx\in U\backslash \{\mathbf{0}\}$ and
\item[(c)]$L^1_{\fb} V(\xx)\leq 0$ for all $\xx\in U$,
\end{enumerate}
then $\mathbf{0}$ is a stable equilibrium of (\ref{eq:ode}). Moreover, if
condition (c) is replaced by
\begin{enumerate}
\item[(c')] $L^1_{\fb} V(\xx)< 0$ for all $\xx\in U\backslash\{\mathbf{0}\}$,
\end{enumerate}
then $\mathbf{0}$ is an asymptotically stable equilibrium of (\ref{eq:ode}).
Such $V$ is called a {\itshape Lyapunov function}.
\end{theorem}
For asymptotic stability, we have
Barbashin-Krasovskii-LaSalle (BKLS) Principle which relaxes condition $(c\textrm{'})$ in Theorem \ref{Lya}.
\begin{theorem}[BKLS Principle] \label{bkls} \
Suppose there exists $V$ satisfying the conditions (a), (b) and (c)
in Theorem \ref{Lya}. If the set $\mathcal M\,\define \, \{\xx\in \mathbb R^n\mid
L^1_{\fb} V(\xx)=0\}\cap U$ does not contain any trajectory of the
system besides the trivial trajectory $\xx(t)\equiv \mathbf{0}$,
then $\mathbf{0}$ is asymptotically stable.
\end{theorem}
Inspired by Theorem \ref{bkls}, we will define \emph{relaxed
Lyapunov function} (RLF for short) in the subsequent section, which
guarantees the asymptotic stability of an equilibrium of a dynamical
system.
\section{Relaxed Lyapunov Function}\label{sec:RLF}
Intuitively, a Lyapunov function requires that any trajectory
starting from $\xx_0\in U$ cannot leave the region $\{\xx\in \mathbb
R^n\mid V(\xx)\leq V(\xx_0)\}$. While, in the asymptotic stability
case, the corresponding $V$ forces any trajectory starting from
$\xx_0\in U$ to transect the boundary $\{\xx\in \mathbb R^n\mid
V(\xx)= V(\xx_0)\}$ towards the set $\{\xx\in \mathbb R^n\mid
V(\xx)< V(\xx_0)\}$. It is clear that $L_{\fb}^1 V(\xx)<0$ is only
a sufficient condition to guarantee asymptotic stability.
When a point $\xx$ satisfies $L_{\fb}^1 V(\xx)=0$,
the transection requirement may still be met if the first non-zero
higher order Lie derivative of $V$ at $\xx$ is negative. To
formalize this idea, we give the following definition.
\begin{definition}[Pointwise Rank]\label{dfn:point-rank} \
Let $\mathbb{N}^+$ be the set of positive natural numbers.
Given sufficiently smooth function $\sigma$ and vector filed $\fb$, the {\itshape pointwise rank} of $\sigma$ w.r.t. $\fb$ is defined as the function $\gamma_{\sigma,\fb}: \mathbb R^n \rightarrow \mathbb N\cup\{\infty\}$ given by
\begin{equation*}
\gamma_{\sigma,\fb}(\xx)= \left\{
\begin{array}{l}
\infty , \qquad \mbox{if } \forall k\in \mathbb{N}^+.\, L^k_{\fb}\sigma(\xx)= 0, \\
\min\{k\in \mathbb{N}^+\mid L^k_{\fb}\sigma(\xx)\neq 0\} , \quad \mbox{otherwise.}
\end{array} \right.
\end{equation*}
\end{definition}
\begin{example}\label{eg:point-rank}
For $\fb=(-x,y)$ and $p(x,y)=x+y^2$, by Example \ref{eg:Lie-derv}, we have $\gamma_{p,\fb}(0,0)=\infty$, $\gamma_{p,\fb}(1,1)=1$, $\gamma_{p,\fb}(2,1)=2$.
\end{example}
\begin{definition}[Transverse Set]\label{dfn:transet}\
Given sufficiently smooth function $\sigma$ and vector field $\fb$, the
\emph{transverse set} of $\sigma$ w.r.t $\fb$ is defined as
\begin{equation*}
\mathrm{Trans}_{\sigma,\fb}\,\define \, \{\xx\in \mathbb R^n \mid \gamma_{\sigma,\fb}(\xx)<\infty \wedge L^{\gamma_{\sigma,\fb}(\xx)}_{\fb} \sigma(\xx)<0\} \enspace .
\end{equation*}
\end{definition}
Intuitively, $\mathrm{Trans}_{\sigma,\fb}$ consists of those points
at which the first non-zero high order Lie derivative of $\sigma$
along $\fb$ is negative. Now we can relax condition $(c\textrm{'})$
in Theorem \ref{Lya} and get a stronger result for asymptotic
stability.
\begin{theorem}\label{LyaSta}\
Suppose $ \mathbf{0} $ is an equilibrium point of
(\ref{eq:ode}). If there is an open set $ U\subset \mathbb R^n$ with
$\mathbf{0}\in U$ and a sufficiently smooth function $V: U\rightarrow \mathbb R$ s.t.
\begin{enumerate}
\item[(a)] $V(\mathbf{0})=0$,
\item[(b)] $V(\xx)>0$ for all $\xx\in U \backslash \{\mathbf{0}\}$ and
\item[(c)] $\xx \in \mathrm{Trans}_{V,\fb}$ for all $\xx\in U \backslash \{\mathbf{0}\}$,
\end{enumerate}
then $\mathbf{0}$ is an asymptotically stable equilibrium
point of (\ref{eq:ode}).
\end{theorem}
\begin{proof}
First notice that condition (c) implies
$L^1_{\fb} V(\xx)\leq 0$ for all $\xx\in U\backslash\{\mathbf{0}\}$.
In order to show the asymptotic stability of $ \mathbf{0}$, according to Theorem \ref{bkls},
it is sufficient to show that $\mathcal M\,\define \, \{\xx \in \mathbb R^n\mid
L^1_{\fb} V(\xx)=0\}\cap U$ contains no
nontrivial trajectory of the dynamical system.
If not, let
$\xx(t)$, $t\geq 0$ be such a trajectory
contained in $\mathcal M$ other than $\xx(t)\equiv \mathbf{0}$. Then $L^1_{\fb}V(\xx(t))=0$ for all $t\geq 0$. Noting that $\xx_0=\xx(0)\in \mathrm{Trans}_{V,\fb}$, we can get
the Taylor expansion of $L^1_{\fb}V(\xx(t))$ at $t=0$:
\begin{align*}
L^1_{\fb}V(\xx(t)) &= L_{\fb}^1 V(\xx_0) + L_{\fb}^2 V(\xx_0)\cdot t +
L_{\fb}^3 V(\xx_0)\cdot \frac{t^2}{2!} +\cdots \\
&= L_{\fb}^{\gamma_{V,\fb}(\xx_0)} V(\xx_0)\cdot
\frac{t^{\scriptscriptstyle{\gamma_{V,\fb}(\xx_0)}-1}}{(\gamma_{V,\fb}(\xx_0)-1)!}+\cdots\enspace .
\end{align*}
By Definition \ref{dfn:transet}, there exists an $\epsilon>0$ s.t. $\forall t \in (0,\epsilon). \,L^1_{\fb}p(\xx(t))<0,$
which contradicts the assumption.
\end{proof}
\begin{definition}[Relaxed Lyapunov Function] \
We refer to the function $V$ in Theorem \ref{LyaSta}
as a {\itshape relaxed Lyapunov function}, denoted by
RLF.
\end{definition}
In the next section, we will explore how to discover polynomial RLFs automatically for PDSs.
\section{Automatically Discovering RLFs for PDSs}\label{sec:main-result}
Given a PDS, the process of automatically discovering polynomial RLFs is as follows:
\begin{itemize}
\item a template, i.e. a parametric polynomial $p(\uu,\xx)$, is predefined as a potential RLF;
\item the conditions for $p(\uu,\xx)$ to be an RLF are translated into an equivalent formula $\Phi$
of
the decidable \emph{first-order theory of reals} \cite{tarski51};
\item constraint $\Phi '$ on parameters $\uu$, or equivalently a set $S_{\uu}$ of all $t$-tuples subject to $\Phi'$, is obtained by applying \emph{quantifier elimination} (QE for short. See \cite{Redlog,qepcad}) to $\Phi$, and any instantiation of $\uu$ by $\uu_0 \in S_{\uu}$ yields an RLF $p_{\uu_0}(\xx)$.
\end{itemize}
\subsection{Computation of Transverse Set}\label{sec:auxiliary}
Correct translation of the three conditions in Theorem \ref{LyaSta} is crucial to our method. In particular, we have to show that for any polynomial $p(\xx)$ and polynomial vector field $\fb$, the transverse set $\mathrm{Trans}_{p,\fb}$ can be represented by first order polynomial formulas. To this end, we first give several theorems by exploring the properties of Lie derivatives and polynomial ideas.
In what follows, given a parameterized polynomial $p(\uu,\xx)$, all Lie derivatives $L_{\fb}^k p$ are seen as polynomials in $\mathbb{R}[\uu,\xx]$. Besides, we will use the convention that $\bigvee_{i\in \emptyset}\eta_i =\textit{false}$ and $\bigwedge_{i\in \emptyset}\eta_i =\textit{true}$, where $\eta_i$ is logical formula.
\begin{theorem}[Fixed Point Theorem]\label{GIT}\
Given $p\,\define\,p(\uu,\xx)$, if $L^{i}_{\fb}p\in \<L^{1}_{\fb}p,$ $ \cdots, L^{i-1}_{\fb}p\rg$,
then for all $m>i$,
$L^{m}_{\fb}p\in \<L^{1}_{\fb}p, \cdots,
L^{i-1}_{\fb}p\rg.$
\end{theorem}
\begin{proof}
We prove this fact by induction. Assume
$L^{k}_{\fb}p\in \<L^{1}_{\fb}p,$ $ \cdots, L^{i-1}_{\fb}p\rg$ for
$k\geq i$. Then there are $g_j\in \mathbb
R[\uu, \xx]$ s.t.
$L^{k}_{\fb}p=\sum_{j=1}^{i-1} g_jL^{j}_{\fb}p$.
By the definition of Lie derivative it follows that
\begin{align}
L^{k+1}_{\fb} p&~=~(\frac{\partial}{\partial \xx}L^k_{\fb}p, \fb) \nonumber\\
~&~=~(\frac{\partial}{\partial \xx}\sum_{j=1}^{i-1} g_jL^{j}_{\fb}p, \fb) \nonumber\\
~&~=~(\sum_{j=1}^{i-1}L^{j}_{\fb} p \frac{\partial}{\partial \xx}g_j+\sum_{j=1}^{i-1}g_j
\frac{\partial}{\partial \xx}L^{j}_{\fb}p, \fb) \nonumber\\
~&~=~\sum_{j=1}^{i-1}(\frac{\partial}{\partial \xx}g_j, \fb)L^{j}_{\fb} p+\sum_{j=1}^{i-1}g_jL^{j+1}_{\fb}p\nonumber\\
~&~=~\sum_{j=1}^{i-1}(\frac{\partial}{\partial \xx}g_j, \fb)L^{j}_{\fb}
p+\sum_{j= 2}^{i-1}g_{j-1}L^{j}_{\fb}p+g_{i-1}L^i_{\fb}p. \nonumber
\end{align}
By induction hypothesis, $L^i_{\fb}p\in \<L^{1}_{\fb}p,
\cdots, L^{i-1}_{\fb}p\rg$, so $L^{k+1}_{\fb}p\in\<L^{1}_{\fb}p,
\cdots, L^{i-1}_{\fb}p\rg$. By induction, the fact follows immediately.
\end{proof}
\begin{theorem}\label{thm:upper}\
Given $p\,\define\,p(\uu,\xx)$, the number
$$N_{p,\fb}=\min\{i\in\mathbb{N}^+\mid L^{i+1}_{\fb}p\in \<L^{1}_{\fb}p, \cdots,
L^{i}_{\fb}p\rg\}$$
is well defined and computable.
\end{theorem}
\begin{proof}
First it is easy to show that $N_{p,\fb}$ has an equivalent expression
$\label{eq:npf}N_{p,\fb}=\min\{i\in\mathbb{N}^+\mid I_{i+1}=I_i\}$,
where $I_i=\<L^{1}_{\fb}p, \cdots,
L^{i}_{\fb}p\rg\subseteq \mathbb R[\uu,\xx]$.
Notice that
$$I_1\subseteq\\I_2\subseteq\cdots\subseteq I_k\cdots $$
forms an ascending chain of ideals. By Theorem \ref{ACC},
$N_{p,\fb}$ is well-defined. Computation of $N_{p,\fb}$ is actually an \emph{ideal membership} problem, which can be solved by computation of \emph{Gr\"{o}bner basis} \cite{clo}.
\end{proof}
\begin{example}
For $\fb=(-x,y)$ and $p(x,y)=x+y^2$, by Example \ref{eg:Lie-derv}, we have $L^2_{\fb}p\notin \<L^1_{\fb}p\rg$ and $L^3_{\fb}p\in \<L^1_{\fb}p,L^2_{\fb}p\rg$, so $N_{p,\fb}=2$\,.
\end{example}
\begin{theorem}[Rank Theorem]\label{Para-Dir-Rank} \ \
Suppose that $p\,\define\, p(\uu,\xx)$. Then for all $\xx\in \mathbb
R^n$ and all $\uu_0\in \mathbb{R}^t$, $\gamma_{p_{\uu_0},\fb}(\xx)<\infty$ {implies} $\gamma_{p_{\uu_0},\fb}(\xx)\leq N_{p,\fb}.$
\end{theorem}
\begin{proof}
If the conclusion is not true, then there exist $\xx_0\in\mathbb R^n$ and $\uu_0\in\mathbb R^t$ s.t.
$$N_{p,\fb}<\gamma_{p_{\uu_0},\fb}(\xx_0)<\infty \enspace .$$
By Definition \ref{dfn:point-rank}, $\xx_0$ satisfies
\begin{equation*}
L_{\fb}^1 p_{\uu_0}=0\wedge \cdots\wedge L_{\fb}^{N_{p,\fb}}p_{\uu_0}=0
\wedge L_{\fb}^{\gamma_{p_{\uu_0},\fb}(\xx_0)}p_{\uu_0}\neq 0\enspace .
\end{equation*}
Then by Theorem \ref{thm:upper} and \ref{GIT},
for all $m>N_{p,\fb}$, we have $L_{\fb}^m p_{\uu_0}(\xx_0)=0$. In particular, $L_{\fb}^{\gamma_{p_{\uu_0},\fb}(\xx_0)}p_{\uu_0}(\xx_0)=0$, which contradicts $L_{\fb}^{\gamma_{p_{\uu_0},\fb}(\xx_0)}p_{\uu_0}(\xx_0)\neq 0$.
\end{proof}
Now we are able to show the computability of $\mathrm{Trans}_{p,\fb}$.
\begin{theorem}\label{thm:comput-trans}\
Given a parameterized polynomial $p\,\define\,p(\uu,\xx)$ and polynomial vector field $\fb$, for any $\uu_0\in \mathbb{R}^t$ and any $\xx\in\mathbb{R}^n$, $\xx\in \mathrm{Trans}_{p_{\uu_0},\fb}$ if and only if $\uu_0$ and $\xx$ satisfy $\varphi_{p,\fb}$, where
\begin{equation}\label{eq:varphi}
\varphi_{p,\fb}\,\define \,
\bigvee_{i=1}^{N_{p,\fb}}\varphi_{p,\fb}^{i}\,,\quad \mbox{and}
\end{equation}
\begin{equation}\label{eq:varphi-i}
\varphi_{p,\fb}^{i}\,\define \, (\bigwedge_{j=1}^{i-1}L_{\fb}^j
p(\uu,\xx)=0)\wedge L_{\fb}^i p(\uu,\xx)<0\,.
\end{equation}
\end{theorem}
\begin{proof}
($\Rightarrow$) Suppose $\xx\in \mathrm{Trans}_{p_{\uu_0},\fb}$. By Definition \ref{dfn:transet} $\xx$ satisfies
\begin{equation}\label{eq:trans2}
L_{\fb}^1 p_{\uu_0}=0\wedge \cdots\wedge L_{\fb}^{\scriptscriptstyle{\gamma_{p_{\uu_0},\fb}(\xx)-1}}p_{\uu_0}=0
\wedge L_{\fb}^{\scriptscriptstyle{\gamma_{p_{\uu_0},\fb}(\xx)}}p_{\uu_0}<0\enspace .
\end{equation}
By Theorem \ref{Para-Dir-Rank}, $\gamma_{p_{\uu_0},\fb}(\xx)\leq N_{p,\fb}$. Then it is easy to check that (\ref{eq:trans2}) implies (\ref{eq:varphi}) when $\uu=\uu_0$.
($\Rightarrow$) If $\uu_0$ and $\xx$ satisfy $\varphi_{p,\fb}$, then from Definition \ref{dfn:transet} we can see that $\xx\in \mathrm{Trans}_{p_{\uu_0},\fb}$ holds trivially.
\end{proof}
\subsection{A Sound and Complete Method for Generating RLFs}
Based on the results established in Section \ref{sec:auxiliary}, we
can give a sound and complete method for automatically generating
polynomial RLFs on PDSs.
Given $\xx =(x_1,x_2,\ldots, x_n)\in \mathbb R^n$, let $\|\xx \|=
\sqrt{\sum_{i=1}^n x_i^2}$ denote the Euclidean norm of $\xx$. Let
$\mathcal B(\xx,d)=\{\mathbf{y}\in \mathbb R^n\mid
\|\mathbf{y}-\xx\|< d\}$ for any $d> 0$. Then our main result can
be stated as follows.
\begin{theorem}[Main Result]\label{thm:main}
Given a PDS $\dot{\xx}=\fb(\xx)$ with $\fb(\mathbf{0})=\mathbf{0}$
and a parametric polynomial $p\,\define\,p(\uu,\xx)$. Let $r_0\in
\mathbb R$ and $\uu_0=(u_{1_0},u_{2_0},\ldots,u_{t_0})\in \mathbb
R^t$. Then $p_{\uu_0}$ is an RLF in $\mathcal {B}(\mathbf{0},r_0)$
if and only if
\[(u_{1_0},u_{2_0},\ldots, u_{t_0},r_0) \in
\mathrm{QE}( \phi_{p,\fb})\enspace ,\] where
\begin{equation}\label{eqn:phi}
\phi_{p,\fb}\,\define \, \phi_{p,\fb}^{1}\wedge
\phi_{p,\fb}^{2} \wedge \phi_{p,\fb}^{3} \enspace ,
\end{equation}
\begin{equation}\label{eq:phi-1}
\phi_{p,\fb}^{1}\,\define \, p(\uu,\mathbf{0})=0\enspace ,
\end{equation}
\begin{equation}\label{eq:phi-2}
\phi_{p,\fb}^{2}\,\define \, \forall \xx.(\| \xx\|^2>0 \, \wedge
\,\| \xx\|^2 < r^2 \rightarrow p(\uu,\xx)>0)\,,
\end{equation}
\begin{equation}\label{eq:phi-3}
\phi_{p,\fb}^{3}\,\define \,\forall \xx.( \| \xx\|^2>0 \, \wedge
\,\| \xx\|^2 < r^2 \rightarrow \varphi_{p,\fb})\enspace .
\end{equation}
\end{theorem}
\begin{proof}
First, in Theorem \ref{LyaSta}, the existence of an open set $U$ is
equivalent to the existence of an open set $\mathcal
{B}(\mathbf{0},r_0)$. Then according to Theorem
\ref{thm:comput-trans}, it is easy to check that (\ref{eq:phi-1}),
(\ref{eq:phi-2}) and (\ref{eq:phi-3}) are direct translations of
conditions $(a)$, $(b)$ and $(c)$ in Theorem \ref{LyaSta}.
\end{proof}
According to Theorem \ref{thm:main}, we can follow the three steps at the beginning of Section \ref{sec:main-result} to discover polynomial RLFs on PDSs. This method is ``complete" because we can discover all possible polynomial RLFs by enumerating all polynomial templates.
\subsection{Implementation}
To construct $\phi_{p,\fb}$ in Theorem \ref{thm:main}, we need to compute $N_{p,\fb}$ in advance, which is time-consuming. What is worse, when $N_{p,\fb}$ is a large number the resulting $\phi_{p,\fb}$ can be a huge formula, for which QE is difficult.
For analysis of asymptotic stability,
one RLF is enough. Therefore if an RLF can be obtained by solving constraint involving merely lower order Lie
derivatives, there's no need to resort to higher order ones. Regarding this, we give an incomplete but more efficient implementation of Theorem \ref{thm:main}, by constructing $\phi_{p,\fb}$ and searching for RLFs in a stepwise manner.
Let
$$
\psi_{p,\fb}^{i}\,\define \, \bigwedge_{j=1}^{i-1} L_{\fb}^j
p(\uu,\xx)=0,\, \mbox{for}\, i\geq 1\enspace ,
$$
\begin{eqnarray*}
\theta_{p,\fb}^{i}\,\define \, \forall \xx.(\|
\xx\|^2>0 \,
\wedge \,\| \xx\|^2 < r^2 \wedge \psi_{p,\fb}^{i}
\rightarrow L_{\fb}^i p(\uu,\xx)<0)
\end{eqnarray*}
and
\begin{eqnarray*}
\bar\theta_{p,\fb}^{i}\,\define \, \forall \xx.(\|
\xx\|^2>0 \,
\wedge \,\| \xx\|^2 < r^2 \wedge \psi_{p,\fb}^{i}
\rightarrow L_{\fb}^i p(\uu,\xx)\leq0)\,.
\end{eqnarray*}
Intuitively, for $\xx$ satisfying $\psi_{p,\fb}^{i}$, we have to
impose constraints $\theta_{p,\fb}^{i}$ or
$\bar\theta_{p,\fb}^{i}$ on the $i$-th higher order Lie derivative
of $p$ along $\fb$.
Now the RLF generation algorithm (RLFG for short) can be formally stated as
follows.
\begin{algorithm}[ht]\label{alg:lfg}
\caption{Relaxed Lyapunov Function Generation}
Input: $\fb \in \mathbb{R}[x_1,\ldots,x_n]^n$
with $\fb(\mathbf{0})=\mathbf{0}$, \\$p\in \mathbb{R}[u_1,\ldots,u_t, x_1,\ldots,x_n] $\\
Output: $Res\subseteq \mathbb{R}^{t+1}$ \\
$i:=1$; $temp:=\emptyset$; $L_{\fb}^1 p:=(\frac{\partial}{\partial \xx} p,\fb)$;\\
$Res^{0}:=\mathrm{QE}(\phi_{p,\fb}^{1} \wedge \phi_{p,\fb}^{2} )$;\\
\If{$Res^{0}=\emptyset$}{return $\emptyset$;}\Else{
\textbf{repeat}\\
\qquad $temp:= Res^{i-1}\cap \mathrm{QE}(\theta_{p,\fb}^{i})$;\\
\qquad \If{$temp \neq \emptyset\,$}{\qquad return $temp\,$;}\qquad
\Else{\qquad $Res^{i}:=Res^{i-1}\cap \mathrm{QE}
(\bar\theta_{p,\fb}^{i})$;\\
\qquad\If{$Res^{i}=\emptyset$}{\qquad return
$\emptyset$;} \qquad \Else{\qquad $i:=i+1$;\\
\qquad $L_{\fb}^i p:=(\frac{\partial}{\partial \xx} L_{\fb}^{i-1}
p,\fb)$;}}\textbf{until}\,\,\,$L_{\fb}^i
p\in \langle L_{\fb}^1 p, L_{\fb}^2 p, \ldots,
L_{\fb}^{i-1}p\rangle$;}
return $\emptyset$;
\end{algorithm}
\begin{remark} Formula $\phi_{p,\fb}^{1}$ and $\phi_{p,\fb}^{2}$ in
line 5 are defined in (\ref{eq:phi-1}) and (\ref{eq:phi-2}); QE in line 5, 10
and 14 is done in a computer algebra tool like Redlog
\cite{Redlog} or QEPCAD \cite{qepcad}; in line 20
the loop test can be done by calling the \emph{IdealMembership} command in Maple$^{\scriptscriptstyle{\textrm{TM}}}$ \cite{Maple}.
\end{remark}
The idea of Algorithm \ref{alg:lfg} is: at the $i$-th step, we
search for an RLF using constraint constructed from Lie derivatives
with order no larger than $i$. If this fails to produce a solution,
then we add the $(i+1)$-th order Lie derivative to the constraint.
This process continues until either we succeed in finding a
solution, or we can conclude that there is no RLF with the
predefined template, or we get to the $N_{p,\fb}$-th iteration,
which means no solution exists at all.
Correctness of the
algorithm RLFG is guaranteed by the following theorem.
\begin{theorem}\label{thm:crrct}
For Algorithm \ref{alg:lfg}, we have
\begin{description}
\item[1) Termination.]\qquad\qquad RLFG terminates for any valid input.
\item[2) Soundness.] \qquad\qquad If $(\uu,r) = \, (u_1,u_2,\ldots,u_t,r)
\in Res$, then $p_{\uu}(\xx)$ is an RLF in
$\mathcal{B}(\mathbf{0},r)$.
\item[3) Weak Completeness.]\qquad\qquad\qquad\qquad If $Res=\emptyset$ then there does
not exist an RLF in the form of $p(\uu,\xx)$.
\end{description}
\end{theorem}
\begin{proof}
\begin{itemize}
\item[(1)] The loop condition is $L_{\fb}^i p \notin \langle L_{\fb}^1 p, L_{\fb}^2 p, \ldots,$ $L_{\fb}^{i-1}p\rangle$. By Theorem \ref{thm:upper}, RLFG can run at most $N_{p,\fb}$
many iterations.
\item[(2)] Suppose $Res^{0},Res^{1},\ldots, Res^{k}$ is the longest sequence generated by RLFG when it terminates. We can inductively prove that this sequence satisfies the following properties.
\begin{itemize}
\item[(P1)] $0\leq k \leq N_{p,\fb}$.
\item[(P2)] $Res^{i}=\mathrm{QE}(\phi_{p,\fb}^{1}\wedge \phi_{p,\fb}^{2} \wedge \bar\phi_{p,\fb}^{i})$, \,\, for $0\leq i \leq k$, where \begin{eqnarray*}\bar\phi_{p,\fb}^{i}\,\define \,\forall \xx. \Big{(}\| \xx\|^2>0 \,
\wedge \,\| \xx\|^2 < r^2)\longrightarrow
\\
\big( (\bigvee_{j=1}^i \varphi_{p,\fb}^{j}\,)
\,\vee \psi_{p,\fb}^{i+1}\big{)}\Big{)}\enspace.
\end{eqnarray*}
\item[(P3)] $\mathrm{QE}(\phi_{p,\fb}^{1}\wedge \phi_{p,\fb}^{2} \wedge \tilde\phi_{p,\fb}^{i})=\emptyset,$ \, for $1\leq i \leq k$, where
$$\tilde\phi_{p,\fb}^{i}\,\define \, \forall \xx. \Big( (\| \xx\|^2>0 \,
\wedge \,\| \xx\|^2 < r^2)
\rightarrow \bigvee_{j=1}^i \varphi_{p,\fb}^{j}\Big)\, .$$
\item[(P4)] $Res=\emptyset $ if and only if either $Res^{k}=\emptyset$ or $k=N_{p,\fb}$; otherwise $Res=\mathrm{QE}( \phi_{p,\fb}^{1}\wedge \phi_{p,\fb}^{2}\wedge \tilde\phi_{p,\fb}^{k+1})$.
\end{itemize}
Suppose $(\uu,r)\in Res$, then by (P1), (P4) and
(\ref{eqn:phi}) we can get $Res \subseteq
\mathrm{QE}({\phi_{p,\fb}})$. Thus $(\uu,r) \in \mathrm{QE}({\phi_{p,\fb}}) $ and $p_{\uu}(\xx)$ is an RLF according to Theorem \ref{thm:main}.
\item[(3)] Suppose $Res=\emptyset$, then by (P4) we have either $k=N_{p,\fb}$ or $Res^{k}=\emptyset$. If $k=N_{p,\fb}(\geq 1)$,
then by (P3) and (\ref{eqn:phi}) we get
$\mathrm{QE}(\phi_{p,\fb})=\emptyset$; if $Res^{k}=\emptyset$, from (P1), (P2), (\ref{eqn:phi}) as well as the validity of
\[
\Big(\bigvee_{j=1}^{N_{p,\fb}} \varphi_{p,\fb}^{j}\Big)
\rightarrow \Big( \big{(}\bigvee_{j=1}^k \varphi_{p,\fb}^{j}\,\big{)}
\,\vee \psi_{p,\fb}^{k+1}\Big), \,\, 0\leq k\leq N_{p, \fb}, \]
we have
$\mathrm{QE}(\phi_{p,\fb})\subseteq Res^{k}=\emptyset$. So far we have proved $Res = \emptyset$ implies $\mathrm{QE}(\phi_{p,\fb})= \emptyset$. Again by applying Theorem \ref{thm:main} we get the final conclusion.
\end{itemize}
\end{proof}
\section{Example}\label{sec:example}
We illustrate our method for RLF generation using the following example.
\newpage
\begin{example}\label{eg:smpl}
Consider the nonlinear dynamical system \begin{equation}\label{eqn:eg1}
\left(\begin{array}{c} \dot x\\ \dot y
\end{array}\right)=\left(\begin{array}{c} -x+y^2\\ -xy
\end{array}\right)
\end{equation}
with a unique equilibrium point $O(0,0)$. We want to establish the asymptotic stability of $O$.
First,
the linearization of (\ref{eqn:eg1}) at $O$ has the coefficient matrix
\begin{displaymath}
A=\left(\begin{array}{cc} -1 & 0\\ 0 & 0 \end{array}\right)
\end{displaymath}
with eigenvalues $-1$ and $0$, so none of the principles of stability for linear systems apply. Besides, a homogeneous quadratic Lyapunov function $x^2+axy+by^2$ for verifying asymptotic stability of (\ref{eqn:eg1}) does not exist in $\mathbb R^2$, because
$$
\mathrm{QE}\Big{(}\forall x\forall y. \left(
\begin{array}{c}
(x^2+y^2> 0\rightarrow x^2+axy+by^2>0)\\
\wedge \,(2x\dot x + ay\dot x +a x\dot y+ 2by\dot y<0)
\end{array}
\right)\Big{)}
$$
is $\textit{false}$.
However, if we try to find an RLF in $\mathbb R^2$ using
the simple template $p\,\define \, x^2+ay^2$, then Algorithm \ref{alg:lfg} returns $a=1$ at the third iteration. This means (\ref{eqn:eg1}) has an RLF $x^2+y^2$, and $O$ is asymptotically stable.
\end{example}
From this example, we can see that RLFs really extend the class of functions that can be used for asymptotic stability analysis, and our method for automatically discovering RLFs can save us a lot of effort in finding conventional Lyapunov functions in some cases.
\section{Conclusions and Future Work}\label{sec:conclusion}
In this paper, we first generalize the notion of Lyapunov functions
to \emph{relaxed Lyapunov functions} by considering the higher order Lie derivatives of a smooth function along a vector field. The main advantage of RLF is that it provides us more probability of certifying asymptotic stability. We also
propose a method for automatically discovering polynomial
RLFs for polynomial dynamical systems. Our method is complete in the sense that we can enumerate all potential polynomial RLFs by enumerating all polynomial templates for a given PDS. We
believe that our methodology could serve as a mathematically
rigorous framework for the asymptotic stability analysis.
The main disadvantage of our approach is the high computational complexity: the complexity of the first-order quantifier elimination over the closed fields of reals is doubly exponential \cite{dh88}.
Currently we are considering improving the efficiency QE on first order polynomial formulas in special forms, and it will be the main focus of our future work.
\section{Acknowledgments}
The authors thank Professors C. Zhou, L. Yang,
B. Xia and W. Yu
for their helpful discussions on the topic.
|
\section{Introduction}
The low energy effective field theory of D-branes consists of the Dirac-Born-Infeld (DBI) \cite{Bachas:1995kx} and the Chern-Simons (CS) actions \cite{Douglas:1995bn}. The curvature corrections to the CS part can be found by requiring that the chiral anomaly on the world volume of intersecting D-branes (I-brane) cancels with the anomalous variation of the CS action. This action for a single D$_p$-brane at order $O(\alpha'^2)$ is given by \cite{Green:1996dd,Cheung:1997az,Minasian:1997mm},
\begin{eqnarray}
S_{_{CS}}&\supset&-\frac{\pi^2\alpha'^2T_{p}}{24}\int_{M^{p+1}}C^{(p-3)}\wedge\bigg[\tr (R_T\wedge R_T)-\tr (R_N\wedge R_N)\bigg]\labell{CS}
\end{eqnarray}
where $M^{p+1}$ represents the world volume of the D$_p$-brane.
For totally-geodesic embeddings of world-volume in the ambient spacetime, ${\rm R}_{T,N}$ are the pulled back curvature 2-forms of the tangent and normal bundles respectively (see the appendix in ref.
\cite{Bachas:1999um} for more details).
The curvature corrections to the DBI action has been found in \cite{Bachas:1999um} by requiring consistency of the effective action with the $O(\alpha'^2)$ terms of the corresponding disk-level scattering amplitude \cite{Garousi:1996ad,Hashimoto:1996kf}. For totally-geodesic embeddings of world-volume in the ambient spacetime, the corrections in the string frame for zero B-field and for constant dilaton are \cite{Bachas:1999um}
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&\frac{\pi^2\alpha'^2T_{p}}{48}\int d^{p+1}x\,e^{-\phi}\sqrt{-G}\bigg[R_{abcd}R^{abcd}-2\hR_{ab}\hR^{ab}-R_{abij}R^{abij}+2\hR_{ij}\hR^{ij}\bigg]\labell{DBI}
\end{eqnarray}
where $\hR_{ab}=G^{cd}R_{cadb}$ and $\hR_{ij}=G^{cd}R_{cidj}$. Here also a tensor with the world-volume or transverse space indices is the pulled back of the corresponding bulk tensor onto world-volume or transverse space\footnote{Our index conversion is that the Greek letters $(\mu,\nu,\cdots)$ are the indices of the space-time coordinates, the Latin letters $(a,d,c,\cdots)$ are the world-volume indices and the letters $(i,j,k,\cdots)$ are the normal bundle indices.}. For the case of D$_3$-brane with trivial normal bundle the curvature couplings \reef{CS} and \reef{DBI} have been modified in \cite{Bachas:1999um} to include the complete sum of D-instanton corrections by requirement of the $SL(2,Z)$ invariance of the couplings.
In the presence of non-constant dilaton, the couplings \reef{DBI} are not consistent with T-duality. For zero B-field, the compatibility with linear T-duality requires the following extension of \reef{DBI}:
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&\frac{\pi^2\alpha'^2T_{p}}{48}\int d^{p+1}x\,e^{-\phi}\sqrt{-G}\bigg[R_{abcd}R^{abcd}-2(\hR_{ab}-\phi_{,ab})(\hR^{ab}-\phi^{,ab})\nonumber\\
&&\qquad\qquad\qquad\qquad\qquad\qquad-R_{abij}R^{abij}+2(\hR_{ij}-\phi_{,ij})(\hR^{ij}-\phi^{,ij})\bigg]\labell{DBI2}
\end{eqnarray}
where commas denote partial differentiation\footnote{Using on-shell relations, the definition of the curvature tensor $\hat{R}_{\mu\nu}$ in \cite{Garousi:2009dj} has been changed as $\hat{R}_{\mu\nu}\equiv \frac{1}{2}(R_{\mu a}{}^a{}_{\nu}-R_{\mu k}{}^k{}_{\nu})$. With this tensor the couplings in \reef{DBI} are then invariant under linear T-duality \cite{Garousi:2009dj} when B-field is zero. If one uses the standard definition $\hat{R}_{\mu\nu}\equiv R_{\mu a}{}^a{}_{\nu}$, then the couplings \reef{DBI2} are invariant.}. The dilaton couplings for D$_3$-brane, however, breaks the S-duality of the curvature terms found in \cite{Bachas:1999um}. Using the fact that the dilaton and the RR zero-form transform similarly under the S-duality transformations, one expects there should be similar couplings as above for the RR zero-form. We will show that the S-matrix element of two RR vertex operators found in \cite{Garousi:1996ad} produces in fact such couplings. Having similar couplings for the dilaton and the RR zero-form, we use an appropriate $SL(2,Z)$ matrix to write the results in $SL(2,R)$ invariant form. However, there is an overall factor of $e^{-\phi}$ in the Einstein frame which is not invariant under the $SL(2,R)$ transformation.
In the presence of non-zero B-field, the couplings \reef{CS} and \reef{DBI2} are not consistent with the T-duality. Using the compatibility of these couplings with linear T-duality as a guiding principle, the quadratic B-field couplings at order $O(\alpha'^2)$ have been found in \cite{Garousi:2009dj,Becker:2010ij,Garousi:2010rn}. The B-field couplings in the DBI part are \cite{Garousi:2009dj}\footnote{In this paper, we are interested only in the quadratic order of field strengths, {\it e.g.,}\ we are not considering $H^4$, $RH^2$, or $R\prt\phi\prt\phi$ terms in DBI action.}
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_{p}}{48}\int d^{p+1}xe^{-\phi}\sqrt{-G}\bigg[\frac{1}{6}H_{ijk,a}H^{ijk,a}+\frac{1}{3}H_{abc,i}H^{abc,i}-\frac{1}{2}H_{bci,a}H^{bci,a}
\bigg]\labell{LDBI}
\end{eqnarray}
The above couplings have been confirmed with the disk level S-matrix calculations \cite{Garousi:2009dj}. These couplings for D$_3$-brane again break the S-duality of the curvature terms found in \cite{Bachas:1999um}. Using the fact that the B-field and the RR two-form appear as a doublet in the S-duality transformations, one expects there should be similar couplings as above for the RR two-form. We will show that the S-matrix element of two RR vertex operators found in \cite{Garousi:1996ad} produces in fact such couplings. Having similar couplings for the B-field and the RR two-form, we then write the results in $SL(2,R)$ invariant form. In this case also, there is an overall factor of $e^{-\phi}$ in the Einstein frame which is not invariant under the $SL(2,R)$ transformation.
It has been shown in \cite{Garousi:2010ki} that the CS part should include couplings which involve linear NSNS field. These couplings have been found by studying the S-matrix element of one RR and one NSNS vertex operators at order $O(\alpha'^2)$ \cite{Garousi:1996ad}. These couplings in string frame are \cite{Garousi:2010ki}\footnote{Using on-shell relations, the definition of the curvature tensor $\hat{R}_{ij}$ in \cite{Garousi:2010ki} has been changed as $\hat{R}_{ij}\equiv \frac{1}{2}(R_{ia}{}^a{}_j-R_{ik}{}^k{}_j)$. With this tensor the coupling $F^{(p+2)}_{a_0\cdots a_pj,i}\hat{R}^{ij}$ is then invariant under linear T-duality \cite{Garousi:2010ki}. If one uses the standard definition $\hat{R}_{ij}\equiv R_{ia}{}^a{}_j$, then the second term in the second line in \reef{LCS} can be written at the linear order as $F^{(p+2)}_{a_0\cdots a_p}{}^{ j,i}(h_{ij,aa}+h_{aa,ij}-h_{ia,aj}-h_{ja,ai}-2\phi_{,ij})/2(p+1)$ where $h$ is the metric perturbation. Under T-duality along the world volume direction $y$, the RR factor $F^{(p+2)}_{a_0\cdots a_p}{}^{ j,i}/(p+1)$ which includes the Killing index $y$, transforms to $F^{(p+1)}_{a_0\cdots a_{p-1}}{}^{ j,i}$. This RR field, however, does not include the Killing index. Hence, the indices $i,j$ in the T-dual theory do not include the Killing index $y$. Using this observation, one can easily verify that the metric/dilaton factor $(h_{ij,aa}+h_{aa,ij}-h_{ia,aj}-h_{ja,ai}-2\phi_{,ij})$ is invariant under the linear T-duality. Hence, the second term in the second line in \reef{LCS} is invariant under the T-duality.}
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_p}{24}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}\left(\frac{1}{2!(p-1)!}[{ F}^{(p)}_{ia_2\cdots a_p,a}H_{a_0a_1}{}^{a,i}-{ F}^{(p)}_{aa_2\cdots a_p,i}H_{a_0a_1}{}^{i,a}]\right.\labell{LCS}\\
&&\left.\qquad\qquad\qquad\qquad+\frac{2}{p!}\bigg[\frac{1}{2!}F^{(p+2)}_{ia_1\cdots a_pj,a}R^a{}_{a_0}{}^{ij}-\frac{1}{p+1}F^{(p+2)}_{a_0\cdots a_pj,i}(\hat{R}^{ij}-\phi^{,ij})\bigg]\right.\nonumber\\
&&\left.\qquad\qquad\qquad\qquad-\frac{1}{3!(p+1)!}F^{(p+4)}_{ia_0\cdots a_pjk,a}H^{ijk,a}\right)\nonumber
\end{eqnarray}
The redundant fields $F^{(6)},\cdots, F^{(9)}$ in above action are related to the magnetic dual of the RR field strengths $F^{(1)},\cdots, F^{(4)}$ as $F^{(10-n)}=*F^{(n)}$ for $n=1,2,3,4$.
For the self-dual D$_3$-brane in the Einstein frame, we will show that up to the overall factor of $e^{-\phi}$, the above couplings can be written in $SL(2,Z)$ invariant forms.
The S-duality of the above couplings is like the S-duality of the $R^4$ corrections to the supergravity action \cite{Green:1981ya,Grisaru:1986dk,Grisaru:1986px,Gross:1986iv}. In the Einstein frame these couplings have an overall factor of $e^{-3\phi/2}$. It has been conjectured in \cite{Green:1997tv} that this factor is in fact the leading order term of the non-holomorphic Eisenstein series $E_s$ with $s=3/2$ at weak coupling. This conjecture has been confirmed with one loop \cite{Green:1997tv} and two loops \cite{D'Hoker:2005jc}. We speculate that the above $O(\alpha'^2)$ corrections to D$_3$-brane action can be written in $SL(2,Z)$ invariant form by extending the weak coupling factor $e^{-\phi}$ to the regularized non-holomorphic Eisenstein series $E_s$ with $s=1$. This function appears also in the $R^2$ terms of the D$_3$-brane action \cite{Bachas:1999um,Basu:2008gt}.
The quadratic B-field couplings have been added to the CS action \reef{CS} by requiring the consistency of this action with the linear T-duality \cite{Becker:2010ij,Garousi:2010rn}. These couplings are
\begin{eqnarray}
S_{_{CS}}&\!\!\!\!\!\supset\!\!\!\!\!&\frac{-\pi^2\alpha'^2T_p}{24\times2!2!(p-3)!}\int d^{p+1}x\epsilon^{a_0\cdots a_{p}}{\cal C}^{(p-3)}_{a_4\cdots a_{p-4}}\bigg[\frac{1}{2}H_{a_0a_1a,i}H_{a_2a_3}{}^{a,i}-\frac{1}{2}H_{a_0a_1i,a}H_{a_2a_3}{}^{i,a}
\bigg]\labell{Tf41new}
\end{eqnarray}
They have been confirmed by the S-matrix element of one RR and two NSNS vertex operators \cite{Becker:2010ij,Garousi:2011ut}. For the D$_3$-brane case, S-duality indicates that there should be similar couplings for $CC^{(2)}C^{(2)}$. Including these couplings one can write them in $SL(2,Z)$ invariant form. However, in this case the overall factor is the axion field which may be written in the S-dual form by replacing it with another $SL(2,Z)$ invariant function $f(\tau,\bar{\tau})$ whose weak-expansion starts at axion instead of the dilaton in the Eisenstein series.
An outline of the paper is as follows: We begin the section 2 by examining the disk-level S-matrix element of two RR vertex operators from which we find the quadratic RR couplings on the world volume of D$_p$-branes at order $O(\alpha'^2)$. In section 3, we show that for the self-dual D$_3$-brane case and for the RR two-form, the couplings are exactly the same as the B-field couplings \reef{LDBI}. We then write the result in $SL(2,Z)$ invariant form using the $SL(2,R)$ matrix $\cM$ which appears in the Type IIB supergravity action, and the regularized non-holomorphic Eisenstein series $E_s$ with $s=1$. In this section we write also the couplings of the RR four-form and the RR scalar in S-dual form using $E_1(\tau,\bar{\tau})$. In section 4, we write the CS couplings in $SL(2,Z)$ invariant form. We end this paper by summarizing the new disk level couplings that are predicted by requring the consistency of the DBI action \reef{DBI} and the CS action \reef{CS} with the S-duality.
\section{RR couplings from S-matrix }
The scattering amplitude of two RR states from D$_p$-brane is given by \cite{Garousi:1996ad}
\begin{eqnarray}
A(\veps_1,p_1;\veps_2,p_2)&=&-\frac{\alpha'^2T_p}{16\times 32}K(1,2)\frac{\Gamma(-\alpha' t/4)\Gamma(\alpha' q^2)}{\Gamma(1-\alpha't/4+\alpha' q^2)}\nonumber\\
&=&\frac{T_p}{16\times 32}K(1,2)\left(\frac{4}{q^2t}+\zeta(2)\alpha'^2+O(\alpha'^4)\right)\labell{Amp}
\end{eqnarray}
where $q^2=p_1^ap_1^b\eta_{ab}$ is the momentum flowing along the world-volume of D-brane, and $t=-(p_1+p_2)^2$ is the momentum transfer in the transverse direction. The kinematic factor is
\begin{eqnarray}
K(1,2)&=&\left(2q^2a_1+\frac{t}{2}a_2\right)\labell{kin}
\end{eqnarray}
where
\begin{eqnarray}
a_1(n,m,p)&=&-\frac{1}{2}\Tr(P_-\Gamma_{1(n)}M_p\gamma_{\mu}C^{-1}M_p^T\Gamma^T_{2(m)}C\gamma^{\mu})
\labell{fintwo}\\
\nonumber\\
a_2(n,m,p)&=&\frac{1}{2}\Tr(P_-\Gamma_{1(n)}M_p\gamma_{\mu})\Tr(P_-\Gamma_{2(m)}M_p\gamma^{\mu})
\labell{finthree}\nonumber
\end{eqnarray}
where
\begin{eqnarray}
\Gamma_{\alpha(n)}=\frac{1}{n!}(F_{\alpha})_{\nu_1\cdots\nu_n}\gamma^{\nu_1}\cdots\gamma^{\nu_n}\nonumber\\
M_p=\frac{\pm 1}{(p+1)!}\epsilon_{a_0\cdots a_p}\gamma^{a_0}\cdots\gamma^{a_p}
\end{eqnarray}
where $F_{\alpha}$ for $\alpha=1,2$ is the linearized RR field strength $n$-form and $\epsilon$ is the volume $p+1$-form of the $D_p$-brane. In equation \reef{kin}, $P_-=\frac{1}{2}(1-\gamma_{11})$ is the chiral projection operator. The $\gamma_{11}$ gives the magnetic couplings and $1$ gives the electric couplings. The first term in \reef{Amp} produces the massless poles resulting from the $(\alpha')^0$ order of the DBI and CS couplings on the D-brane, and the supergravity couplings in the bulk. The second term in \reef{Amp} should produce $(\alpha')^2$ couplings of two RR fields on the world volume of D$_p$-brane in which we are interested.
Using various identities, $a_1$ can be simplified for electric components of the RR field strength to \cite{Garousi:1996ad}
\begin{eqnarray}
a_1(n,m,p)&=&\frac{8}{n!}\delta_{mn}\bigg[\Tr(D)F_{1(n)}\!\cdot\! F_{2(n)}-2nD^{\lambda}{}_{\kappa}F_{1\lambda\nu_2\cdots \nu_n} F_2^{\kappa\nu_2\cdots \nu_n}\bigg]
\end{eqnarray}
where the matrix $D^{\mu}_{\nu}$ is diagonal with +1 in the world volume directions and -1 in the transverse directions. The degree of the RR field strength $n$ in $a_1$ is independent of the dimension of the D$_p$-brane. For the magnetic components, one finds the same result but for different $n$, {\it i.e.,}\ $n'=10-n$. Using the fact that the redundant field strength $F^{(10-n)}$ for $n=1,2,3,4$ are the magnetic dual of $F^{(n)}$, {\it i.e.,}\ $F^{(10-n)}=*F^{(n)}$, one finds the following result for the magnetic components of the RR vertex operator:
\begin{eqnarray}
a_1(n',m',p)&=&\frac{8}{n'!}\delta_{m'n'}\bigg[\Tr(D)(*F_{1(n)})\!\cdot\! (*F_{2(n)})-2n'D^{\lambda}{}_{\kappa}(*F^{(n)})_{1\lambda\nu_2\cdots \nu_{n'}} (*F^{(n)})_2^{\kappa\nu_2\cdots \nu_{n'}}\bigg]\nonumber
\end{eqnarray}
The degree of the RR field strength in $a_2$ depends on the dimension of the D$_p$-brane. To see this consider the factor $\Tr(P_-\Gamma_{1(n)}M_p\gamma_{\mu})$ in $a_2$. For the electric components it is nonzero only for $n=p$ and for $n=p+2$, and for magnetic components it is nonzero for $(10-n)=p$ and for $(10-n)=p+2$. Performing the trace in each case one finds for the electric components
\begin{eqnarray}
\Tr(P_-\Gamma_{1(p)}M_p\gamma_{\mu})&=&\frac{16}{p!}\delta^{a_0}{}_{\mu}F_1^{a_1\cdots a_p}\epsilon_{a_0\cdots a_p}\nonumber\\
\Tr(P_-\Gamma_{1(p+2)}M_p\gamma_{\mu})&=&\frac{16}{(p+1)!}F_1^{a_0\cdots a_p}{}_{\mu}\epsilon_{a_0\cdots a_p}
\end{eqnarray}
We have not pay attention to the signs on the right hand sides because we are interested in $a_2$ which is quadratic multiple of each term. One can easily verify that $a_2$ is nonzero only for the cases
\begin{eqnarray}
a_2(n=m=p)&=&\frac{8\times 16}{n!}F^{(n)}_{1}\!\cdot\! V\!\cdot\! F^{(n)}_{2}\nonumber\\
a_2(n=m=p+2)&=&\frac{8\times 16}{(n-1)!}(F_1^{(n)})^{a_0\cdots a_p}{}_{i}(F_2^{(n)})_{a_0\cdots a_p}{}^{i}\nonumber
\end{eqnarray}
where the notation $F^{(n)}_{1}\!\cdot\! V\!\cdot\! F^{(n)}_{2}$ means the indices are contracted with the world volume metric $\eta^{ab}$.
For the magnetic components, one finds
\begin{eqnarray}
a_2(n'=m'=p)&=&\frac{8\times 16}{n'!}(*F_1^{(n)})\!\cdot\! V\!\cdot\! (*F_2^{(n)})\nonumber\\
a_2(n'=m'=p+2)&=&\frac{8\times 16}{(n'-1)!}(*F_1^{(n)})^{a_0\cdots a_p}{}_{i}(*F_2^{(n)})_{a_0\cdots a_p}{}^{i}\nonumber
\end{eqnarray}
where $n'=10-n$. Again the kinematic factor $a_2$ for the magnetic components is the same as for the electric components but for $*F$.
The kinematic factor for the electric RR fields is
\begin{eqnarray}
K(1,2)&=&-\sum_{n}\left(2a_1p_1\!\cdot\! V\!\cdot\! p_2+a_2p_1\!\cdot\! p_2\big[\delta_{n,p}+\delta_{n,p+2}\big]\right)
\end{eqnarray}
where the summation is over $n=1,2,3,4,5$. The field theory corresponding to this kinematic factor for D$_p$-brane in the string frame is
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_p}{48\times 32}\int d^{p+1}x\,e^{\phi}\sqrt{-G}\sum_n\frac{16}{n!}\bigg[(p-4)F^{(n)}{}_{,a}\!\cdot\! F^{(n)\,,a}-nD^{\mu}{}_{\nu}F^{(n)}{}_{\mu}{}_{,a}\!\cdot\! F^{(n)\nu}{}^{,a}\nonumber\\
&&+4F^{(n)}{}_{,\mu}\!\cdot\! V\!\cdot\! F^{(n)\,,\mu}\delta_{n,p}+4nF^{(n)}{}_{i,\mu}\!\cdot\! V\!\cdot\! F^{(n)\,i,\mu}\delta_{n,p+2}\bigg]\labell{FF}
\end{eqnarray}
where we have also used the standard convention that the RR fields are rescaled as $C\rightarrow e^{\phi}C$. This is the reason why there is the dilaton factor $e^{\phi}$ in above action instead of the expected factor of $e^{-\phi}$ for the disk amplitude. Similar rescaling has been also used in couplings \reef{LCS}.
The magnetic couplings are the same as above with replacing $F$s with $*F$s.
The above action gives all quadratic RR couplings at order $O(\alpha'^2)$ for any D$_p$-brane. However, we are interested in the world volume couplings of the self-dual D$_3$-brane.
\section{S-duality of DBI part}
Let us start with the couplings of the two-forms. The couplings of NSNS two-form are given in \reef{LDBI} which have been found in \cite{Garousi:2009dj} by consistency of the curvature couplings with T-duality. Compatibility of these couplings with S-duality requires similar couplings for the RR two-form. Writing the spacetime indices in \reef{FF} in terms of the world volume and the transverse indices, one finds that for D$_3$-brane and for RR two-form the above action simplifies to
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_{3}}{48}\int d^{4}x\,e^{\phi}\sqrt{-G}\bigg[\frac{1}{6}F_{ijk,a}F^{ijk,a}+\frac{1}{3}F_{abc,i}F^{abc,i}-\frac{1}{2}F_{bci,a}F^{bci,a}
\bigg]\labell{FF1}
\end{eqnarray}
These couplings are similar to the B-field couplings in \reef{LDBI}.
To study the transformation of the above couplings under S-duality, one should first rescale the metric from string frame to the Einstein frame $G_{\mu\nu}=e^{\phi/2}g_{\mu\nu}$. The B-field couplings \reef{LDBI} are multiplied by $e^{-2\phi}$ and the dilaton drops out of the above RR couplings. The D$_3$-brane and the Einstein frame metric are invariant under $SL(2,Z)$, and the B-field and the RR two-form transform as doublet, {\it i.e.,}\
\begin{eqnarray}
{\cal B}^{(2)}=\pmatrix{B^{(2)} \cr
C^{(2)}}
\end{eqnarray}
Under a $SL(2,Z)$ transformation by
\begin{eqnarray}
\Lambda=\pmatrix{d&c \cr
b&a}
\end{eqnarray}
the ${\cal B}$-field transforms linearly by the rule
\begin{eqnarray}
{\cal B}\rightarrow \Lambda {\cal B}
\end{eqnarray}
The axion and the dilaton combine into a complex scalar field $\tau=C+ie^{-\phi}$. This field transforms as
\begin{eqnarray}
\tau\rightarrow \frac{a\tau+b}{c\tau +d}
\end{eqnarray}
Now consider the matrix ${\cal M}$
\begin{eqnarray}
{\cal M}=e^{\phi}\pmatrix{|\tau|^2&-C \cr
-C&1}\labell{M}
\end{eqnarray}
which transforms under the $SL(2,Z)$ as
\begin{eqnarray}
{\cal M}\rightarrow (\Lambda ^{-1})^T{\cal M}\Lambda ^{-1}
\end{eqnarray}
This matrix appears in the $SL(2,Z)$ form of the type IIB supergravity.
Using this matrix, one can rewrite the couplings \reef{LDBI} and \reef{FF1} in the Einstein frame as:
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\!\!\supset\!\!\!\!\!\!&-\frac{\pi^2\alpha'^2T_{3}}{48}\int d^{4}x\,e^{-\phi}\sqrt{-g}\bigg[\frac{1}{6}{\cal F}^T_{ijk,a}\cM\cF^{ijk,a}+\frac{1}{3}\cF^T_{abc,i}\cM\cF^{abc,i}-\frac{1}{2}\cF^T_{bci,a}\cM\cF^{bci,a}
\bigg]\labell{FF2}
\end{eqnarray}
where $\cF=d\cB$. Apart from the overall dilaton factor $e^{-\phi}$, it is invariant under the $SL(2,Z)$ transformation.
The above situation is like the $R^4$ corrections to the supergravity action which apart from the factor $\zeta(3)e^{-3\phi/2}$, the couplings are invariant under the $SL(2,Z)$ transformation. In that case, the tree level, one loop and one-instanton results are the three leading order terms of the non-holomorphic Eisenstein series $E_s$ with $s=3/2$. So it has been conjectured in \cite{Green:1997tv} that the $SL(2,Z)$ invariant coupling is $E_{3/2}R^4$. In particular, this conjecture indicates that there is no perturbative corrections to $R^4$ other than one loop. It has been shown in \cite{D'Hoker:2005jc} that there is no two loop correction to this action.
For general $s$, this series is defined by
\begin{eqnarray}
2\zeta(2s)E_s(\tau,\bar{\tau})&=&\sum_{(m,n)\neq(0,0)}\frac{\tau_2^s}{|m+n\tau|^{2s}}\labell{series}
\end{eqnarray}
where $\tau=\tau_1+i\tau_2$. It is invariant under the $SL(2,Z)$ transformation. This function satisfies the following differential equation:
\begin{eqnarray}
4\tau_2^2\prt_{\tau}\prt_{\bar{\tau}}E_s&=&s(s-1)E_s
\end{eqnarray}
This equation has two solutions $\tau_2^s$ and $\tau_2^{1-s}$ corresponding to two particular orders of perturbation theory, and infinite number of non-perturbative solutions. For $s=1$, however, the series \reef{series} diverges logarithmically. The regularized function has the following expansion \cite{Bachas:1999um,Basu:2008gt}:
\begin{eqnarray}
2\zeta(2)E_1(\tau,\bar{\tau})&=&\zeta(2)\tau_2-\frac{\pi}{2}\ln(\tau_2)+\pi\sqrt{\tau_2}\sum_{m\neq 0,n\neq 0}\left|\frac{m}{n}\right|^{1/2}K_{1/2}(2\pi|mn|\tau_2)e^{2\pi imn\tau_1}\labell{regE}
\end{eqnarray}
where the first term corresponds to $n=0$ in the series \reef{series}. This term is exactly the dilaton factor in \reef{FF2}. This may indicate that the factor $e^{-\phi}$ in \reef{FF2} should be replaced by the $SL(,Z)$ invariant function $E_1$.
Another evidence for this replacement is the following. The B-field couplings \reef{LDBI} are related to the gravity couplings \reef{DBI} by T-duality \cite{Garousi:2009dj}. On the other hand the $SL(2,Z)$ invariant form of the gravity couplings has this nonperturbative factor \cite{Bachas:1999um,Basu:2008gt}. In fact, the regularized function \reef{regE} is proportional to $\log(\tau_2|\eta(\tau)|^4)$ \cite{Bachas:1999um} where $\eta(\tau)$ is the Dedekind $\eta$-function. The $\log\tau_2$ piece is nonanalytic and comes from the annulus \cite{Basu:2008gt} and the remaining part appears in the Wilsonian effective action found in \cite{Bachas:1999um}. Hence, one expects the $SL(2,Z)$ invariant form of the $\cB$-field couplings \reef{FF2} to be
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_{3}}{24}\int d^{4}x\,E_1(\tau,\bar{\tau})\sqrt{-g}\bigg[\frac{1}{6}{\cal F}^T_{ijk,a}\cM\cF^{ijk,a}\nonumber\\
&&\qquad\qquad\qquad+\frac{1}{3}\cF^T_{abc,i}\cM\cF^{abc,i}-\frac{1}{2}\cF^T_{bci,a}\cM\cF^{bci,a}
\bigg]\labell{FF3}
\end{eqnarray}
The second term in the expansion of $E_1$ is the annulus contribution to the 1PI effective action. All other terms are D-instanton contributions.
The next case that we consider is the couplings of two $F^{(5)}$ on the world volume of D$_3$-brane. The RR potential $C^{(4)}$ is invariant under the $SL(2,Z)$ transformation. One can easily confirm that the couplings \reef{FF} for $C^{(4)}$ in the Einstein frame have the dilaton factor $e^{-\phi}$. Replacing this factor with the $SL(2,Z)$ invariant function $E_1(\tau,\bar{\tau})$, one finds the following S-dual couplings:
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_3}{48\times 5!}\int d^{4}x\,E_1(\tau,\bar{\tau})\sqrt{-g}\bigg[-F^{(5)}{}_{,a}\!\cdot\! F^{(5)\,,a}-5D^{\mu}{}_{\nu}F^{(5)}{}_{\mu}{}_{,a}\!\cdot\! F^{(5)\nu}{}^{,a}\nonumber\\
&&+20F^{(5)}{}_{i,\mu}\!\cdot\! V\!\cdot\! F^{(5)\,i,\mu}\bigg]\labell{FF5}
\end{eqnarray}
where we have also included the $*F^{(5)}$ terms. Similar couplings as those in the first line above, without the factor $E_1$, can be written for the $F^{(5)}$ couplings on the world volume of D$_7$-brane. Since it is S-duality invariant, the $F^{(5)}$ couplings in \reef{FF} are the couplings for all $(p,q)$ 7-branes.
The disk level S-matrix element of two NSNS vertex operators shows that there is no couplings for one dilaton and one graviton at order $O(\alpha'^2)$ for D$_3$-brane \cite{Garousi:2009dj}. The kinematic factor for D$_3$-brane is
\begin{eqnarray}
K(\phi, \veps)&=&\Tr(\veps)\left(\frac{4q^4}{\sqrt{8}}\right)\labell{phih}
\end{eqnarray}
where $\veps$ is the polarization of the symmetric tensor. Using the traceless of the graviton polarization, one finds the above kinematic factor is zero for graviton\footnote{This term which is zero for one graviton and one dilaton, did not considered in \cite{Garousi:2009dj}. This term has non-zero contribution for the kinematic factor of two diatons.}. This is consistent with S-duality because RR couplings \reef{LCS} have no coupling between one graviton and one axion. It is also consistent with the couplings \reef{DBI2}. Using various on-shell relations, one can show that in the Einstein frame these couplings become
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&\frac{\pi^2\alpha'^2T_{p}}{48}\int d^{p+1}x\,e^{-\phi}\sqrt{-g}\bigg[R_{abcd}R^{abcd}-2\hR_{ab}\hR^{ab}
-R_{abij}R^{abij}+2\hR_{ij}\hR^{ij}\nonumber\\
&&-(p-3)\big[\hR_{ab}\phi^{,ab}-\hR_{ij}\phi^{,ij}\big]-\frac{(p-3)^2}{8}\big[\phi_{,ab}\phi^{,ab}-\phi_{,ij}\phi^{,ij}]+\phi_{,ab}\phi^{,ab}\bigg]\labell{DBI3}
\end{eqnarray}
which gives zero coupling for the linear dilaton for the case $p=3$. Replacing the polarization tensor $\veps$ in \reef{phih} with the dilaton polarization, one finds the last term in the above equation.
Under the S-duality both the dilaton and the RR scalar appears in the complex field $\tau$, so one expects that there should be similar coupling for the RR scalar at order $O(\alpha'^2)$ for D$_3$-brane. The couplings \reef{FF} for the RR scalar for the D$_3$-brane in the Einstein frame become
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&\frac{\pi^2\alpha'^2T_3}{48}\int d^{4}x\,e^{\phi}\sqrt{-g}C_{,ab}C^{,ab}\labell{FF6}
\end{eqnarray}
To study the S-duality of this coupling we add the dilaton coupling in \reef{DBI3} to the above equation, {\it i.e.,}\
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&\frac{\pi^2\alpha'^2T_3}{48}\int d^{4}x\,e^{-\phi}\sqrt{-g}\bigg[\phi_{,ab}\phi^{,ab}+e^{2\phi}C_{,ab}C^{,ab}\bigg]\labell{FF611}
\end{eqnarray}
The terms in the bracket can be extended to the $SL(2,Z)$ invariant form by including the disk-level nonlinear terms, and the dilaton factor can be extended to the $SL(2,Z)$ invariant form by including the annulus and the D-instanton effects.
The S-dual extension of the coupling \reef{FF611} is
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\!\supset\!\!\!\!&-\frac{\pi^2\alpha'^2T_3}{96}\int d^{4}x\,E_1(\tau,\bar{\tau})\sqrt{-g}\Tr(\cM_{,ab}(\cM^{-1})^{,ab})\labell{FF61}
\end{eqnarray}
where the matrix $\cM$ is the one appears in \reef{M}
\section{S-duality of CS part}
The linear NSNS couplings in the CS action are given in \reef{LCS}. To study the S-duality of these couplings for the self-dual D$_3$-brane, we begin by witting the couplings in the first line in the Einstein frame, {\it i.e.,}\
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{24\times 2!2!}\int d^{4}x\,\epsilon^{a_0\cdots a_3}e^{-\phi}\bigg[{ F}_{ia_2a_3,a}H_{a_0a_1}{}^{a,i}-{ F}_{aa_2 a_3,i}H_{a_0a_1}{}^{i,a}\bigg]\labell{LCS1}
\end{eqnarray}
To write it in a $SL(2,Z)$ invariant form, we first introduce the $SL(2,Z)$ matrix
\begin{eqnarray}
\cN=\pmatrix{0&1 \cr
-1&0}
\end{eqnarray}
This matrix has the property
\begin{eqnarray}
\cN=(\Lambda ^{-1})^T{\cal N}\Lambda ^{-1}
\end{eqnarray}
Using this matrix, one can rewrite the terms in the parenthesis in the S-dual form $\cF^T\cN\cF$. Then using the same idea as in previous section, one can extend it to the following $SL(2,Z)$ invariant form:
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{12\times 2!2!}\int d^{4}x\,\epsilon^{a_0\cdots a_3}E_1(\tau,\bar{\tau}){ \cF^T}_{a_0 a_1a,i}\cN \cF_{a_2a_3}{}^{i,a}\labell{LCS2}
\end{eqnarray}
The second term in the expansion of $E_1(\tau,\bar{\tau})$ can be calculated with the annulus level S-matrix element of one RR and one NSNS vertex operators.
The RR couplings in the second line of \reef{LCS} for the D$_3$-brane in the Einstein frame are proportional to the dilaton factor $e^{-\phi}$. In particular, the term $ \hat{\cR}^{ij}-\phi^{,ij}$ becomes proportional to $\hat{\cR}^{ij}+\eta^{ij}\phi_{,a}{}^a/4$ in the Einstein frame. The last term gives $F^{(5)}_{a_0\cdots a_3i}{}^{,i}$ which is zero on-shell, {\it i.e.,}\ $F^{(5)}_{a_0\cdots a_3i}{}^{,i}=-F^{(5)}_{a_0\cdots a_3a}{}^{,a}$ where two world volume indices in $F^{(5)}_{a_0\cdots a_3a}$ is identical, hence, it is zero. Using the fact that the Einstein frame metric and $C^{(4)}$ are invariant under the S-duality, and replacing $e^{-\phi}$ with $E_1(\tau,\bar{\tau})$, one finds the following S-dual couplings:
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{6}\int d^{4}x\,\epsilon^{a_0\cdots a_3}E_1(\tau,\bar{\tau})\bigg[\frac{1}{2!3!}F^{(5)}_{ia_1\cdots a_3j,a}\cR^a{}_{a_0}{}^{ij}-\frac{1}{4!}F^{(5)}_{a_0\cdots a_3j,i}\hat{\cR}^{ij}\bigg]\nonumber
\end{eqnarray}
Note that the above couplings are not invariant under the standard form of T-duality transformations. This is because the standard T-duality rules are in string frame whereas the above S-dual couplings are in the Einstein frame. To study the T-duality of the above couplings one should first write the T-duality transformations in the Einstein frame and then applied to the above action.
The coupling in the third line of \reef{LCS} for D$_3$-brane is
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&\frac{\pi^2\alpha'^2T_3}{24}\int d^{4}x\,\epsilon^{a_0\cdots a_3}\left(\frac{1}{3!4!}F^{(7)}_{ia_0\cdots a_3jk,a}H^{ijk,a}\right)\labell{FH}
\end{eqnarray}
Using the fact that the redundant RR field strength $F^{(7)}$ is the magnetic dual of $F^{(3)}$, {\it i.e.,}\ $F^{(7)}=*F^{(3)}$, and the relation
\begin{eqnarray}
(*F^{(n)})_{\mu_{n+1}\cdots \mu_{10}}=\frac{1}{n!}\sqrt{-G}\epsilon_{\mu_1\cdots \mu_{10}}(F^{(n)})^{\mu_1\cdots \mu_n}
\end{eqnarray}
one finds that in the Einstein frame again the coupling \reef{FH} has the dilaton factor $e^{-\phi}$. Moreover, the coupling is antisymmetric under changing the RR two form with the B-field. So the coupling \reef{FH} can be written in the Einstein frame as
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&\frac{\pi^2\alpha'^2T_3}{24\times 2}\int d^{4}x\,\epsilon^{a_0\cdots a_3}e^{-\phi}\left(\frac{1}{3!4!}(*F^{(3)})_{ia_0\cdots a_3jk,a}H^{ijk,a}-\frac{1}{3!4!}(*H)_{ia_0\cdots a_3jk,a}(F^{(3)})^{ijk,a}\right)\nonumber
\end{eqnarray}
Defining the magnetic dual of ${\cal F}^{(3)}=d\cB^{(2)}$ as
\begin{eqnarray}
{\cal F}^{(7)}=*{\cal F}^{(3)}=\pmatrix{*H^{(3)} \cr
*F^{(3)}}\, ,
\end{eqnarray}
one finds the following S-dual coupling:
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{24\times 3!4!}\int d^{4}x\,\epsilon^{a_0\cdots a_3}E_1(\tau,\bar{\tau})({\cal F}^{(7)})^T_{a_0\cdots a_3ijk,a}\cN {\cal F}^{ijk,a}\labell{FH2}
\end{eqnarray}
Similar coupling as above, without the factor $E_1$, can be written for the couplings in the first line of \reef{LCS} for D$_7$-brane. Therefore, the couplings in the first line of \reef{LCS} are the couplings for all $(p,q)$ 7-branes.
Finally we consider the Chern-Simons couplings. Compatibility of the CS action \reef{CS} with T-duality requires the couplings \reef{Tf41new}. We now try to write these B-field couplings in S-dual invariant form. The couplings \reef{Tf41new} for the self-dual D$_3$-brane are
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{48\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}} C\bigg[H_{a_0a_1a,i}H_{a_2a_3}{}^{a,i}-H_{a_0a_1i,a}H_{a_2a_3}{}^{i,a}
\bigg]\labell{HH}
\end{eqnarray}
The above couplings have been also confirmed with scattering calculation \cite{Becker:2010ij,Garousi:2011ut}. The compatibility of these couplings with S-duality indicates that the disk level S-matrix element of three RR vertex operators produces the following couplings:
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{48\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}}e^{2\phi} C\bigg[F_{a_0a_1a,i}F_{a_2a_3}{}^{a,i}-F_{a_0a_1i,a}F_{a_2a_3}{}^{i,a}
\bigg]\labell{FF0}
\end{eqnarray}
where we have also used the standard rescaling $C\rightarrow e^{\phi}C$. In the Einstein frame the terms in the brackets can be combined into the $SL(2,Z)$ invariant form of $\cF^T\cM\cF$. However, in this case amplitude has no overall dilaton factor $e^{-\phi}$. Instead, it has the RR scalar as the overall factor.
The S-dual invariant of this coupling then does not include the non-holomorphic Eisenstein function. It should include another modular function $f(\tau,\bar{\tau})$, which has the following weak-expansion:
\begin{eqnarray}
f(\tau,\bar{\tau})&\sim&\tau_1+\cdots\labell{ff}
\end{eqnarray}
where dots stands for loops and D-instanton effects.
It has been shown in \cite{Bachas:1999um} that the gravity couplings in the effective Wilsonian CS theory for trivial normal bundle can be written in S-dual invariant form. The modular function that appears in this case is $\log(\eta(\tau)/\eta(\bar{\tau}))$. This function has the following weak-expansion \cite{Bachas:1999um}:
\begin{eqnarray}
\log\frac{\eta(\tau)}{\eta(\bar{\tau})}&=&\frac{i\pi}{6}\tau_1-\bigg[q+\frac{3}{2}q^2+\frac{4}{3}q^3+\cdots -cc\bigg]
\end{eqnarray}
where $q=e^{2\pi i\tau}$. The first term arises from the disk level amplitude and the series of the power $q$ stand for the D-instanton corrections. The annulus effect is absent in the above function as it does not contribute to the Wilsonian effective action. On the other hand the function $\log(\eta(\tau)/\eta(\bar{\tau}))$ is not invariant under the $SL(2,Z)$ transformation. The variation of the Wilsonian effective action cancels the anomalous contributions of the massless modes of D$_3$-brane \cite{Bachas:1999um}. The annulus effect in the 1PI effective action should make this function to be $SL(2,Z)$ invariant. So we expect the $SL(2,Z)$ invariant function $f(\tau,\bar{\tau})$ to be
\begin{eqnarray}
f(\tau,\bar{\tau})&=&\log\frac{A(\tau)\eta(\tau)}{A(\bar{\tau})\eta(\bar{\tau})}\labell{f}
\end{eqnarray}
where $A(\tau)$ is the annulus effect which makes $f(\tau,\bar{\tau})$ to be $SL(2,Z)$ invariant.
In terms of this function,
one can write the above couplings in the following S-dual form:
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&\frac{i\pi\alpha'^2T_3}{8\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}} f(\tau,\bar{\tau})\bigg[\cF^T_{a_0a_1a,i}\cM\cF_{a_2a_3}{}^{a,i}-\cF^T_{a_0a_1i,a}\cM\cF_{a_2a_3}{}^{i,a}
\bigg]\labell{FF11}
\end{eqnarray}
It would be interesting to perform the annulus calculation to find the function $A(\tau)$.
For D$_7$-brane, the couplings are the same as above with replacing $f(\tau,\bar{\tau})$ with the $SL(2,Z)$ invariant field $C^{(4)}$. The couplings \reef{HH} have been verified by the S-matrix element of one RR and two NSNS vertex operators \cite{Becker:2010ij,Garousi:2011ut}. The S-matrix element produces some other couplings as well \cite{Garousi:2011ut}. Using the same steps as for \reef{HH}, one can write them in $SL(2,Z)$ invariant forms. This ends our illustration of the consistency of the D-brane action at order $O(\alpha'^2)$ with S-duality.
The consistency of the DBI action \reef{DBI} with S-duality predicts the following couplings in the Einstein frame at the disk level:
\begin{eqnarray}
S_{_{DBI}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_{3}}{48}\int d^{4}x\sqrt{-g}C^2\bigg[\frac{1}{6} H_{ijk,a}H^{ijk,a}+\frac{1}{3}H_{abc,i}H^{abc,i}-\frac{1}{2}H_{bci,a}H^{bci,a}
\bigg]\\
S_{_{DBI}}&\!\!\!\!\!\!\supset\!\!\!\!\!\!&\frac{\pi^2\alpha'^2T_{3}}{24}\int d^{4}x\sqrt{-g}C\bigg[\frac{1}{6} H_{ijk,a}F^{ijk,a}+\frac{1}{3}H_{abc,i}F^{abc,i}-\frac{1}{2}H_{bci,a}F^{bci,a}\nonumber
\bigg]
\end{eqnarray}
The consistency of the CS action \reef{CS} with S-duality predicts the couplings \reef{FF0}.
\begin{eqnarray}
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{48\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}}e^{\phi} C^3\bigg[H_{a_0a_1a,i}H_{a_2a_3}{}^{a,i}-H_{a_0a_1i,a}H_{a_2a_3}{}^{i,a}
\bigg]\nonumber\\
S_{_{CS}}&\!\!\!\supset\!\!\!&\frac{\pi^2\alpha'^2T_3}{24\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}}e^{\phi} C^2\bigg[H_{a_0a_1a,i}F_{a_2a_3}{}^{a,i}-H_{a_0a_1i,a}F_{a_2a_3}{}^{i,a}
\bigg]\\
S_{_{CS}}&\!\!\!\supset\!\!\!&-\frac{\pi^2\alpha'^2T_3}{48\times 2!2!}\int d^{4}x\epsilon^{a_0\cdots a_{3}}e^{\phi} C\bigg[F_{a_0a_1a,i}F_{a_2a_3}{}^{a,i}-F_{a_0a_1i,a}F_{a_2a_3}{}^{i,a}
\bigg]\nonumber
\end{eqnarray}
Moreover, since the modular functions $E_1(\tau,\bar{\tau})$ and $f(\tau,\bar{\tau})$ have annulus contributions, the S-duality also predicts the annulus level couplings for the S-duality invariant couplings that we have found, {\it e.g.,}\ \reef{FF3}. It would be interesting to confirm these couplings by direct calculations.
{\bf Acknowledgments}: I would like to thank Michael B. Green for useful discussions. This work is supported by Ferdowsi University of Mashhad under grant p/17212(1389/12/24).
\bibliographystyle{/Users/Nick/utphys}
\bibliographystyle{utphys} |
\section*{Introduction}
Enumeration of families of plane maps, that is, plane embeddings of
graphs, has received a lot of attention since the 60's; several
methods can be applied: the recursive method introduced by
Tutte~\cite{Tu63}, the random matrix method introduced by Br\'ezin et
al~\cite{Bre}, and the bijective method introduced by Cori and
Vauquelin~\cite{CoriVa} and Schaeffer~\cite{S-these}. In the first
part of this note, we show another method for the enumeration of
rooted simple quadrangulations and triangulations based on quotienting
symmetric simple versions of them. Historically, the enumeration of
symmetric maps of order $k$ (\emph{i.e.}\xspace, such that a rotation of order $k$
fixes the map) was reduced to the enumeration of rooted maps via a
quotient argument, a method used by Liskovets~\cite{Li78}. We proceed
in the reverse way, namely we use two quotient operations on symmetric
simple quadrangulations and triangulations to build in each case an
algebraico-differential equation (Equations~\eqref{eq:q}
and~\eqref{eq:t}) satisfied by the generating series of rooted
corresponding simple maps, which can be explicitly solved to obtain
the formulas for the number of rooted simple quadrangulations (due to
Tutte~\cite{Tu63} and bijectively proved by Schaeffer~\cite{S-these})
and of rooted simple triangulations (due to Tutte~\cite{Tu62} and
bijectively proved by Poulalhon and Schaeffer~\cite{PoSc06}). One
quotient operation is classical and is decribed in
Section~\ref{sec:clasquo}; the other quotient operation is new and, as
described in Section~\ref{sec:ori}, relies deeply on the existence and
properties of $\alpha$-orientations; the new equations for generating
series of simple quadrangulations and triangulations are derived and
solved in Section~\ref{sec:equation}.
The results in the second part are expressions of the series of
several families of symmetric quadrangular and triangular dissections
with control on the distance from the central vertex to the outer
boundary. We recall that symmetric dissections have been counted
according to the number of inner faces by
Brown~\cite{Brown_trig,Brown_quad} using the recursive method
(Liskovet's quotient method~\cite{Li78} can also be applied, reducing
the enumeration to rooted quadrangular dissections). Our approach,
developed in Sections~\ref{sec:dist} and~\ref{sec:dist_trig}, relies
on the quotient method and substitution operations combined with
results by Bouttier \emph{et al.}\xspace~\cite{BoDFGu03,BoGu12}, which express the
series of quadrangulations or triangulations with a marked vertex and
marked edge at prescribed distance from each other.
Our expressions illustrate again the property that the series
expression of a ``well behaved'' map family \ensuremath{\mathcal{M}}\xspace refined by a distance
parameter $d$ is typically expressed in terms of the $d$th power of an
algebraic series of singularity type $z^{1/4}$ (implying that
asymptotically the distance parameter $d$ on a random map of size $n$
in \ensuremath{\mathcal{M}}\xspace converges in the scale $n^{1/4}$ as a random variable).
\section{Plane maps, symmetry and classical quotient}\label{sec:clasquo}
\subsection{Triangulations, quadrangulations and dissections}
A \emph{plane map} is a connected graph embedded in the plane up to
continuous deformation; the unique unbounded face of a plane map is
called the \emph{outer face}, the other ones are called \emph{inner
faces}. Vertices and edges are also called outer if they belong to
the outer face and inner otherwise. A map is said to be \emph{rooted}
if an edge of the outer face is marked and oriented so as to have the
outer face on its left. This edge is the \emph{root edge}, and its
origin is the \emph{root vertex}. A map is \emph{pointed} if one of
its \emph{inner} vertices is marked. For any map \scalebox{.8}{\ensuremath{M}}\xspace, we denote by
$\ensuremath{\mathcal{V}}\xspace(\scalebox{.8}{\ensuremath{M}}\xspace)$, $\ensuremath{\mathcal{F}}\xspace(\scalebox{.8}{\ensuremath{M}}\xspace)$, $\ensuremath{\mathcal{E}}\xspace(\scalebox{.8}{\ensuremath{M}}\xspace)$ its sets of vertices, faces and
edges, and by $v(\scalebox{.8}{\ensuremath{M}}\xspace)$, $f(\scalebox{.8}{\ensuremath{M}}\xspace)$, $e(\scalebox{.8}{\ensuremath{M}}\xspace)$ their cardinalities.
Triangulations and quadrangulations are respectively maps with all
faces of degree 3 or 4, and to avoid the degenerated cases, the outer
face is required to be a simple cycle. For $k\geqslant 1$ and $d\geq3$, a
\emph{$d$-angular dissection of a $k$-gon} or \emph{$d$-angular
$k$-dissection} is a map whose outer face contour is a simple cycle
of length $k$, and with all inner faces of one and the same
degree~$d$. A dissection is said to be \emph{triangular} if $d$
equals~3, \emph{quadrangular} if $d$ equals~4. Observe that
quadrangular $k$-dissections can only exist for even $k$.
A map is said to be \emph{simple} if it has no multiple edges; a
$d$-angular $k$-dissection is called \emph{irreducible} if the
interior of every cycle of length at most $d$ is a face.
\subsection{Symmetric maps and classical quotient}\label{sub:clasquo}
For $k\geqslant 2$, a dissection \scalebox{.8}{\ensuremath{D}}\xspace is said to be \emph{$k$-symmetric} if
its plane embedding (conveniently deformed) is invariant by a
$2\pi/k$-rotation centered at a vertex -- called the \emph{center}
of~\scalebox{.8}{\ensuremath{D}}\xspace.
As observed by Liskovets~\cite{Li78}, any two semi-infinite straight
lines starting from the center and forming an angle of $2\pi/k$
delimit a sector of~\scalebox{.8}{\ensuremath{D}}\xspace. When keeping only this sector and pasting
these two lines together, we obtain a plane map, called the
\emph{$k$-quotient map} of \scalebox{.8}{\ensuremath{D}}\xspace. In other words, the $2\pi/k$-rotation
defines equivalence relations on the sets $\ensuremath{\mathcal{V}}\xspace(\scalebox{.8}{\ensuremath{D}}\xspace)$ and $\ensuremath{\mathcal{E}}\xspace(\scalebox{.8}{\ensuremath{D}}\xspace)$,
and the quotient map of \scalebox{.8}{\ensuremath{D}}\xspace is the map in which equivalent vertices
and equivalent edges are
identified. Figure~\ref{fig:classical_quotient} shows the example of
two symmetric dissections of an hexagon and their quotients. Denote
by $o(D)$ the degree of the outer face of a dissection. The following
lemma is straightforward:
\begin{figure}
\centering
\subfigure[\label{fig:3-sym}]{\includegraphics[page=1,scale=0.7]{hexagone_sym}}\quad
\subfigure[\label{fig:3-quo}]{\includegraphics[page=2,scale=0.7]{hexagone_sym}}\qquad\vline\qquad
\subfigure[\label{fig:2-sym}]{\includegraphics[page=3,scale=0.7]{hexagone_sym}}\qquad
\subfigure[\label{fig:2-quo}]{\includegraphics[page=4,scale=0.7]{hexagone_sym}}
\caption{Examples of symmetric 6-dissections. A 3-symmetric
quadrangular dissection~\subref{fig:3-sym}. Its
3-quotient~\subref{fig:3-quo} is a pointed quadrangular 2-dissection.
A 2-symmetric triangular dissection~\subref{fig:2-sym}. Its
2-quotient~\subref{fig:2-quo} is a pointed triangulation.}
\label{fig:classical_quotient}
\end{figure}
\begin{lemma}\label{lem:quotient_count}
For $k\geqslant 2$, let $\scalebox{.8}{\ensuremath{D}}\xspace$ be a $k$-symmetric dissection,
and $\scalebox{.8}{\ensuremath{E}}\xspace$ its $k$-quotient; we have:
\[
v(\scalebox{.8}{\ensuremath{D}}\xspace) - 1 = k\, (v(\scalebox{.8}{\ensuremath{E}}\xspace)-1) , \quad e(\scalebox{.8}{\ensuremath{D}}\xspace) = k\,e(\scalebox{.8}{\ensuremath{E}}\xspace), \quad f(\scalebox{.8}{\ensuremath{D}}\xspace)
- 1 = k\, (f(\scalebox{.8}{\ensuremath{E}}\xspace)-1), \quad o(\scalebox{.8}{\ensuremath{D}}\xspace)=k\,o(\scalebox{.8}{\ensuremath{E}}\xspace).
\]
\end{lemma}
The quotient operation clearly preserves the degrees of vertices and
faces, hence quotients of $d$-angular dissections are $d$-angular
dissections. More precisely, as any $k$-symmetric dissection is
implicitly pointed (at the center), its $k$-quotient is a pointed
dissection. We define the \emph{radial distance} $r(\scalebox{.8}{\ensuremath{D}}\xspace)$ of a
pointed dissection \scalebox{.8}{\ensuremath{D}}\xspace as the distance between its marked vertex and
the outer face boundary (for instance, the example of
Figure~\ref{fig:3-quo} has radial distance~2 while that of
Figure~\ref{fig:2-quo} has radial distance~1). We also $\ell(\scalebox{.8}{\ensuremath{D}}\xspace)$ the
length of a shortest cycle strictly enclosing $u$.
The following lemma
summarizes the relations between distances in $k$-symmetric
dissections and their $k$-quotients:
\begin{lemma}\label{lem:quotient_distance}
For $k\geqslant 2$, let $\scalebox{.8}{\ensuremath{D}}\xspace$ be a $k$-symmetric dissection,
and $\scalebox{.8}{\ensuremath{E}}\xspace$ its $k$-quotient; we have:
\[
r(\scalebox{.8}{\ensuremath{D}}\xspace)=r(\scalebox{.8}{\ensuremath{E}}\xspace) \quad\text{and}\quad \ell(\scalebox{.8}{\ensuremath{D}}\xspace)=k\,\ell(\scalebox{.8}{\ensuremath{E}}\xspace).
\]
\end{lemma}
\begin{proof}
Let \scalebox{.8}{\ensuremath{P}}\xspace be a pointed dissection, with pointed vertex $u$. A
labelling-function is a function $\lambda: \ensuremath{\mathcal{V}}\xspace(\scalebox{.8}{\ensuremath{P}}\xspace) \to\mathbb{Z}$
such that $\lambda(u)=0$ and for all adjacent vertices $v$ and $v'$,
$|\lambda(v)-\lambda(v')|\leqslant 1$. It is easy to check that the
function $\delta_{P}$ giving the distance from $u$, called
distance-labelling, is the unique labelling-function such that each
vertex $v\neq u$ has a neighbour of smaller label. With this
characterization, it is straightforward that $\delta_{E}$ is the
quotient of $\delta_{D}$; in particular, the radial distance is
the same in \scalebox{.8}{\ensuremath{D}}\xspace as in~\scalebox{.8}{\ensuremath{E}}\xspace.
The second statement is equivalent to the following one: one among
the cycles of minimal length stricly enclosing the center is itself
symmetric. Indeed, let \scalebox{.8}{\ensuremath{C}}\xspace be a cycle of minimal length $\ell(\scalebox{.8}{\ensuremath{D}}\xspace)$
strictly enclosing the center that is minimal for inclusion, and
suppose that \scalebox{.8}{\ensuremath{C}}\xspace is asymmetric. Let \scalebox{.8}{\ensuremath{\tilde C}}\xspace be its image by the
$\pi/k$-rotation; as they both enclose the center, \scalebox{.8}{\ensuremath{C}}\xspace and \scalebox{.8}{\ensuremath{\tilde C}}\xspace are
intersecting cycles and hence define two other cycles enclosing the
center with total length $2 \ell(\scalebox{.8}{\ensuremath{D}}\xspace)$, hence each of length
$\ell(\scalebox{.8}{\ensuremath{D}}\xspace)$; one of these two cycles is included in \scalebox{.8}{\ensuremath{C}}\xspace and \scalebox{.8}{\ensuremath{\tilde C}}\xspace,
contradiction.
\end{proof}
\subsection{Symmetric simple quadrangulations and triangulations
A pointed dissection is called \emph{quasi-simple} if the pointed
vertex lies strictly in the interior of every 1-cycle or 2-cycle; in
particular each vertex can carry at most one loop, otherwise the two
loops would form a 2-cycle that does not contain the pointed vertex.
Unlike that of simple dissections, the family of quasi-simple
dissections is stable under quotient:
\begin{lemma}\label{lem:quotient_quasisimple}
For $k\geqslant 2$, let $\scalebox{.8}{\ensuremath{D}}\xspace$ be a $k$-symmetric dissection and $\scalebox{.8}{\ensuremath{E}}\xspace$ its
$k$-quotient -- both canonically pointed. Then $\scalebox{.8}{\ensuremath{D}}\xspace$ is
quasi-simple if and only if $\scalebox{.8}{\ensuremath{E}}\xspace$ is quasi-simple.
\end{lemma}
\begin{proof}
This just follows from the fact that any 2-cycle of $\scalebox{.8}{\ensuremath{D}}\xspace$ not
enclosing the pointed vertex yields a 2-cycle of $\scalebox{.8}{\ensuremath{E}}\xspace$ not enclosing
the pointed vertex, and reciprocally.
\end{proof}
Note also that, according to Lemma~\ref{lem:quotient_distance}, no
cycle enclosing the center of a $k$-symmetric dissection \scalebox{.8}{\ensuremath{D}}\xspace can be of
length smaller than~$k$. In particular, \scalebox{.8}{\ensuremath{D}}\xspace has no loop, and no
2-cycle enclosing the center if $k>2$. Hence if $k>2$, $k$-symmetric
quasi-simple dissections are indeed simple.
\paragraph{Symmetric simple quadrangulations}
Lemma~\ref{lem:quotient_count} implies that quadrangulations, having
outer degree~4, may only be 2- or 4-symmetric. Moreover, their
quotients are quadrangular dissections, hence they are bipartite,
which prevents them of containing loops, in particular the outer face
has length at least~2. Hence quadrangulations may only be 2-symmetric
(which we simply call \emph{symmetric}), and quotients of symmetric
quadrangulations are pointed quadrangular 2-dissections. This implies
also that quasi-simple symmetric quadrangulations are indeed
simple. Hence the following proposition is a direct consequence of
Lemma~\ref{lem:quotient_quasisimple}:
\begin{proposition}\label{prop:clasquoquad}
The 2-quotient is a one-to-one correspondence between symmetric
simple quadrangulations with $2n$ inner faces and quasi-simple
pointed 2-dissections with $n$ inner faces.
\end{proposition}
\paragraph{Symmetric simple triangulations}
Similarly, triangulations may only be 3-symmetric (which we simply
call \emph{symmetric}), and the 3-quotient of a symmetric
triangulation is a pointed triangular 1-dissection. Note that a
quasi-simple symmetric triangulation is simple (indeed, by
Lemma~\ref{lem:quotient_distance}, a 3-symmetric dissection $\scalebox{.8}{\ensuremath{D}}\xspace$
satisfies $\ell(\scalebox{.8}{\ensuremath{D}}\xspace)\geqslant 3$). Hence:
\begin{proposition}\label{prop:quoclastrig}
The 3-quotient is a one-to-one correspondence between symmetric
simple triangulations with $3n$ inner faces and quasi-simple pointed
triangular 1-dissections with $n$ triangular faces, where $n$ is any
positive odd integer.
\end{proposition}
\section{Another quotient for symmetric simple quadrangulations and
triangulations}\label{sec:ori}
In this section, we only consider simple quadrangulations and
triangulations, and show that a specific quotient can be defined for
these families, using their characterization in terms of
$d$-orientations.
\subsection{The minimal 2- or 3-orientation}\label{sub:ori}
An \emph{orientation} of a plane map is the choice of an orientation
for each of its \emph{inner} edges. An orientation without
counterclockwise cycle is said to be \emph{minimal}. A $d$-orientation of a
map is an orientation such that the outdegree of each inner vertex is
equal to $d$, while outer vertices have outdegree equal to zero.
An important property of simple quadrangulations and triangulations is
that they can be characterized in terms of $d$-orientations:
\begin{proposition}[\cite{Schnyder},\cite{DeOs01},\cite{Felsner2004}]
A quadrangulation is simple if and only if it admits a
2-orientation. A triangulation is simple if and only if it admits a
3-orientation. In both cases, there exists a unique minimal such
orientation.
\end{proposition}
Figure~\ref{fig:symquad} and~\ref{fig:symtrig} show two examples of a
simple quadrangulation and a simple triangulation respectively endowed
with their minimal 2- or 3-orientation.
Now let \scalebox{.8}{\ensuremath{M}}\xspace be a $k$-symmetric map endowed with a $d$-orientation \scalebox{.8}{\ensuremath{O}}\xspace;
then the image of \scalebox{.8}{\ensuremath{O}}\xspace by the $2\pi/k$-rotation is clearly a
$d$-orientation, and it is minimal if \scalebox{.8}{\ensuremath{O}}\xspace is minimal. Hence:
\begin{lemma}
The unique minimal 2-orientation (\emph{resp.}\xspace 3-orientation) of a
symmetric simple quadrangulation (\emph{resp.}\xspace triangulation) is itself
symmetric.
\end{lemma}
\subsection{Leftmost paths}
Let \scalebox{.8}{\ensuremath{M}}\xspace be a plane map endowed with an orientation. For each inner
vertex $u$ of \scalebox{.8}{\ensuremath{M}}\xspace, a \emph{leftmost path starting at $u$} is a maximal
oriented path \scalebox{.8}{\ensuremath{P}}\xspace starting at $u$ such that for any triple $v,v',v''$
of successive vertices along \scalebox{.8}{\ensuremath{P}}\xspace, $(v',v'')$ is the first outgoing
edge after $(v,v')$ in clockwise order around $v'$.
\begin{lemma}\label{lem:leftmostpath}
If \scalebox{.8}{\ensuremath{M}}\xspace is a simple quadrangulation or triangulation endowed with its
minimal 2- or 3- orientation, \scalebox{.8}{\ensuremath{P}}\xspace is necessarily a simple path
ending at one outer vertex of~\scalebox{.8}{\ensuremath{M}}\xspace.
\end{lemma}
\begin{proof}
Assume by contradiction that \scalebox{.8}{\ensuremath{P}}\xspace is self intersecting and let
$v_0,v_1,\ldots, v_l=v$ be a sequence of vertices of \scalebox{.8}{\ensuremath{P}}\xspace explored
successively in this order and forming a cycle. Since the
orientation \scalebox{.8}{\ensuremath{O}}\xspace of \scalebox{.8}{\ensuremath{M}}\xspace is chosen to be minimal, this cycle is
necessarily clockwise. Hence, since \scalebox{.8}{\ensuremath{P}}\xspace is a left-most path, the
outgoing edges of each vertex of the cycle lie in the interior
region of the cycle (except eventually the other outgoing edges of
$v_0$, if $v_0$ is the starting point of the path), hence denoting
by \scalebox{.8}{\ensuremath{R}}\xspace the enclosed region, $e(\scalebox{.8}{\ensuremath{R}}\xspace)$ is at least $2v(\scalebox{.8}{\ensuremath{R}}\xspace)-1$ (\emph{resp.}\xspace
$3v(\scalebox{.8}{\ensuremath{R}}\xspace)-2$). Euler relation leads to a contradiction: if the length
of the cycle is $\ell$, Euler formula implies that $2v(\scalebox{.8}{\ensuremath{R}}\xspace) = e(\scalebox{.8}{\ensuremath{R}}\xspace)
+ 2 + \ell/2$ (\emph{resp.}\xspace $3v(\scalebox{.8}{\ensuremath{R}}\xspace) = e(\scalebox{.8}{\ensuremath{R}}\xspace) + 2 + \ell$). Note that this
is in accordance with the well-known property that inner edges of a
simple triangulation (\emph{resp.}\xspace quadrangulation) can be partitionned
into three (\emph{resp.}\xspace two) spanning trees (see~\cite{Schnyder}).
\end{proof}
In the next two subsections, we use leftmost paths starting at the
center to decompose symmetric simple quadrangulations and
triangulations in 2 (\emph{resp.}\xspace 3) sectors, and show how this particular
sector can be sticked together again in a non-classical way to obtain
a simple quadrangulation or triangulation.
\subsection{A new way of quotienting symmetric simple quadrangulations}
\begin{figure}[t]
\centering
\def.25\linewidth{.25\linewidth}
\def12em{10em}
\begin{minipage}[c]{.3\linewidth} \centering
\subfigure[\scriptsize Symmetric simple quadrangulation]{\qquad
\includegraphics[height=12em, page=1]{Quotient}\quad
\label{fig:symquad}
}
\end{minipage}
\begin{minipage}[c]{3em}\centering
$\nearrow$
\vskip5em
$\searrow$
\end{minipage}
\begin{minipage}[c]{.25\linewidth}\centering
\subfigure[\scriptsize Classical identification]{
\includegraphics[height=12em, page=3]{Quotient}
\label{fig:ident-classique}
}\\
\subfigure[\scriptsize New identification]{
\includegraphics[height=12em, page=2]{Quotient}
\label{fig:idnew}
}
\end{minipage}
\begin{minipage}[c]{3em}\centering
$\longrightarrow$
\vskip10em
$\longrightarrow$
\end{minipage}
\begin{minipage}[c]{.25\linewidth}\centering
\subfigure[\scriptsize Classical quotient]{
\includegraphics[height=12em, page=5]{Quotient}
\label{fig:quoclassique}
}\\
\subfigure[\scriptsize New quotient]{
\includegraphics[height=12em,page=4]{Quotient}
\label{fig:quonew}
}
\end{minipage}
\label{fig:subfigureExample}
\caption{Classical 2-quotient
\subref{fig:ident-classique}, \subref{fig:quoclassique} and the
new quotient \subref{fig:idnew}, \subref{fig:quonew} of a symmetric simple quadrangulation endowed
with its minimal 2-orientation \subref{fig:symquad}.}
\end{figure}
Let \scalebox{.8}{\ensuremath{Q}}\xspace be a symmetric simple quadrangulation, $u$ the central vertex,
$e_1$, $e_2$ its two outgoing edges, and $P_1=(u=v_0,v_1,\ldots,v_p)$,
$P_2=(u=w_0,w_1,\ldots,w_p)$ the leftmost paths of $e_1$ and $e_2$
respectively. Clearly $P_1$ and $P_2$ map to one another by the
$\pi$-rotation. Hence $P_1$ and $P_2$ cannot meet except at their
starting point $u$, otherwise they would meet twice, which would imply
the existence of an oriented cycle without outgoing edges, leading to
the same contradiction as in Lemma~\ref{lem:leftmostpath}.
Let us cut \scalebox{.8}{\ensuremath{Q}}\xspace along $P_1\cup P_2$ to split \scalebox{.8}{\ensuremath{Q}}\xspace into two isomorphic
dissections, see Figure~\ref{fig:idnew}, and define $\scalebox{.8}{\ensuremath{Q}}\xspace_1$ as the one
with clockwise contour $u, v_1, v_2,\dots, w_1$. If $\scalebox{.8}{\ensuremath{Q}}\xspace_1$ is a
quadrangulation, we set $\Phi(\scalebox{.8}{\ensuremath{Q}}\xspace):=\scalebox{.8}{\ensuremath{Q}}\xspace_1$ and mark the edge
$(u,v_1)$. Otherwise, for any $i\leqslant p-2$, we identify in $\scalebox{.8}{\ensuremath{Q}}\xspace_1$
vertices $v_{i+2}$ with $w_i$, and merge corresponding edges; this
defines the map $\Phi(\scalebox{.8}{\ensuremath{Q}}\xspace)$, in which we then mark the edge $(v_1,v_2)$.
Concerning orientations, the identification of $v_{i+2}$ with $w_i$
creates an orientation conflict only when merging $(u,v_1)$ with
$(v_1,v_2)$. We choose to orient the merged edge from $v_1$ to $v_2$.
With this convention, $\Phi(\scalebox{.8}{\ensuremath{Q}}\xspace)$ is naturally endowed with its minimal
$2$-orientation and the leftmost path of the marked edge $(v_1,v_2)$
is $(v_1,v_2,\ldots,v_p)$ (to justify that the $2$-orientation of $\Phi(\scalebox{.8}{\ensuremath{Q}}\xspace)$ thus obtained
is minimal, one just has to observe that no oriented path goes from some vertex $w_i$
to some vertex $v_j$, hence when doing the identifications of vertices
no counterclockwise circuit is created).
It is then easy to describe the inverse
mapping
and to obtain:
\begin{theorem}\label{th:bij}
The mapping $\Phi$ is a one-to-one correspondence between symmetric
simple quadrangulations with $2n$ inner faces and simple
quadrangulations with $n$ inner faces and a marked edge.
\end{theorem}
\subsection{A new way of quotienting symmetric simple triangulations}
\begin{figure}[t]
\centering
\def12em{12em}
\begin{minipage}[c]{.32\linewidth}
\subfigure[\scriptsize Symmetric simple triangulation]{\quad
\includegraphics[height=12em,page=1]{QuotientTrig.pdf}\quad
\label{fig:symtrig}
}
\end{minipage}
\begin{minipage}[c]{2em}\centering
$\nearrow$ \vskip5em $\searrow$
\end{minipage}
\begin{minipage}[c]{.25\linewidth}
\subfigure[\scriptsize Classical identification]{\quad
\includegraphics[height=12em,page=3]{QuotientTrig.pdf}
\label{fig:ident-classique_trig}\hspace{-2em}
}\\
\subfigure[\scriptsize New identification]{\quad
\includegraphics[height=12em,page=2]{QuotientTrig.pdf}\
\label{fig:idnew_trig}\hspace{-2em}
}
\end{minipage}
\begin{minipage}[c]{2em}\centering
$\longrightarrow$ \vskip10em $\longrightarrow$
\end{minipage}\hspace{-1em}
\begin{minipage}[c]{.25\linewidth}
\subfigure[\scriptsize Classical 3-quotient]{
\includegraphics[height=12em,page=5]{QuotientTrig.pdf}
\label{fig:quoclassique_trig}
}\\
\subfigure[\scriptsize New quotient]{\qquad
\includegraphics[height=12em,page=4]{QuotientTrig.pdf}
\label{fig:quonew_trig}
}
\end{minipage}
\label{fig:subfigureExample_trig}
\caption{Classical 3-quotient
\subref{fig:ident-classique_trig}, \subref{fig:quoclassique_trig} and the
new quotient \subref{fig:idnew_trig}, \subref{fig:quonew_trig} of a symmetric
simple triangulation endowed
with its minimal 3-orientation \subref{fig:symtrig}.}
\end{figure}
Let \scalebox{.8}{\ensuremath{T}}\xspace be a simple symmetric triangulation endowed with its minimal
3-orientation. Let $u$ be its central vertex, $e_1$, $e_2$ and $e_3$
its three outgoing edges, and $\scalebox{.8}{\ensuremath{P}}\xspace_1=(u=v_0,v_1,\ldots,v_p)$,
$\scalebox{.8}{\ensuremath{P}}\xspace_2=(u=w_0,w_1,\ldots,w_p)$ and $\scalebox{.8}{\ensuremath{P}}\xspace_3(u=x_0,x_1,\ldots,x_p)$ the
leftmost paths of $e_1$, $e_2$ and $e_3$ respectively. Clearly $\scalebox{.8}{\ensuremath{P}}\xspace_1$,
$\scalebox{.8}{\ensuremath{P}}\xspace_2$ and $\scalebox{.8}{\ensuremath{P}}\xspace_3$ maps respectively onto $\scalebox{.8}{\ensuremath{P}}\xspace_2$, $\scalebox{.8}{\ensuremath{P}}\xspace_3$ and $\scalebox{.8}{\ensuremath{P}}\xspace_1$ by the
$2\pi/3$-rotation centered at $u$. Hence they cannot meet
except at their starting point $u$. Let us cut \scalebox{.8}{\ensuremath{T}}\xspace along $\scalebox{.8}{\ensuremath{P}}\xspace_1\cup
\scalebox{.8}{\ensuremath{P}}\xspace_2\cup \scalebox{.8}{\ensuremath{P}}\xspace_3$ to split \scalebox{.8}{\ensuremath{T}}\xspace into three isomorphic dissections, see
Figure~\ref{fig:idnew_trig}, and define $\scalebox{.8}{\ensuremath{T}}\xspace_1$ as the one with clockwise
contour $u, v_1, v_2,\dots, w_1$. If $\scalebox{.8}{\ensuremath{T}}\xspace_1$ is a triangulation, we set
$\ensuremath{\Phi_{\!\small\triangle}\!}(\scalebox{.8}{\ensuremath{T}}\xspace):=\scalebox{.8}{\ensuremath{T}}\xspace_1$ and mark the edge $(u,v_1)$. Otherwise, for any $i\leqslant
p-2$, we identify in $\scalebox{.8}{\ensuremath{T}}\xspace_1$ vertices $v_{i+2}$ with $w_i$, and merge
corresponding edges; this defines the map $\ensuremath{\Phi_{\!\small\triangle}\!}(\scalebox{.8}{\ensuremath{T}}\xspace)$, in which we then
mark the edge $(v_1,v_2)$.
Concerning orientations, the identification of $v_{i+2}$ with $w_i$
creates an orientation conflict only when merging $(u,v_1)$ with
$(v_1,v_2)$. We choose to orient the merged edge from $v_1$ to $v_2$.
With this convention, $\Phi(Q)$ is naturally endowed with its minimal
$3$-orientation and the leftmost path of the marked edge $(v_1,v_2)$
is $(v_1,v_2,\ldots,v_p)$. It is then easy to describe the inverse
mapping and to obtain:
\begin{theorem}\label{thm:bijtrig}
The mapping $\ensuremath{\Phi_{\!\small\triangle}\!} $ is a one-to-one correspondence between
symmetric simple triangulations with $3n$ inner faces and simple
triangulations with $n$ inner faces and one marked edge, for any
positive odd integer~$n$.
\end{theorem}
\section{Simple quadrangulations and triangulations via symmetric
ones}\label{sec:equation}
\subsection{Getting a functional equation for simple quadrangulations}
Proposition~\ref{prop:clasquoquad} and Theorem~\ref{th:bij} describe
two bijections between symmetric simple quadrangulations and two other
families that are hence also in one-to-one correspondence:
quasi-simple pointed 2-dissections and simple quadrangulations with a
marked edge. Because 2-dissections are bipartite, each pointed
2-dissection corresponds to two different rooted pointed
2-dissections. Similarly, any edge of a simple quadrangulation has an
implicit orientation given by its minimal 2-orientation, hence a
simple quadrangulation with a marked edge corresponds to two distincts
quadrangulations with a marked oriented edge, that can be seen as
rooted simple quadrangulations with a marked face (possibly the outer
one). Hence we obtain:
\begin{corollary}\label{cor}
Rooted simple quadrangulations with $n$ inner faces and a marked
face are in one-to-one correspondence with rooted quasi-simple
pointed 2-dissections with $n$ inner faces.
\end{corollary}
This correspondence allows us to get a functional equation for the
generating series of the family \ensuremath{\mathcal{Q}}\xspace of non-degenerated (meaning, with at least $2$ faces)
rooted simple
quadrangulations. Let $q(x)=\sum_{n\geqslant 2}q_nx^n$ be the series of \ensuremath{\mathcal{Q}}\xspace
according to the number of faces (including the outer one). The
generating series of rooted simple quadrangulations with a marked face
is then equal to $q'(x)$. We want to express also the family $\ensuremath{\mathcal{D}}\xspace_4$ of
rooted quasi-simple 2-dissections in terms of~\ensuremath{\mathcal{Q}}\xspace.
\begin{figure}
\centering
\subfigure[Nested
2-cycles]{\includegraphics[scale=.9, page=3]{decomp_quad}}
\qquad\quad
\subfigure[Element of
$\mathcal{A}^{\scriptscriptstyle{\blacksquare}}$]{\includegraphics[scale=.9,
page=1]{decomp_quad}}
\qquad\quad
\subfigure[Element of
$\mathcal{A}^{\scriptscriptstyle{-}}$]{\includegraphics[scale=.9, page=2]{decomp_quad}}
\caption{The decomposition of a quasi-simple quadrangular 2-dissection.}
\label{fig:decas}
\end{figure}
Observe that for any $\scalebox{.8}{\ensuremath{D}}\xspace\in \ensuremath{\mathcal{D}}\xspace_4$, its separating 2-cycles are
nested and therefore ordered from innermost to outermost. This yields
a decomposition of $\scalebox{.8}{\ensuremath{D}}\xspace_4$ as a sequence of components. Let \ensuremath{\mathcal{A}^\carre}\xspace (\emph{resp.}\xspace
\ensuremath{\mathcal{A}^\trait}\xspace) be the family of rooted 2-dissections with a marked inner vertex
(\emph{resp.}\xspace with a marked inner edge) and where the unique 2-cycle is the
outer face contour. We can cut \scalebox{.8}{\ensuremath{D}}\xspace along the nested 2-cycles to
obtain a pointed 2-dissection in \ensuremath{\mathcal{A}^\carre}\xspace and a sequence of maps in \ensuremath{\mathcal{A}^\trait}\xspace
(see Figure~\ref{fig:decas}). Denoting respectively by $d(x)$,
$a^{\scalebox{.4}{\ensuremath{\blacksquare}}\xspace}(x)$, $a^-\xspace(x)$ the generating series of \ensuremath{\mathcal{D}}\xspace, \ensuremath{\mathcal{A}^\carre}\xspace, and
\ensuremath{\mathcal{A}^\trait}\xspace according to the number of quadrangular faces, this gives:
\begin{equation}\label{eq:d}
d(x)=\frac{a^{\scalebox{.4}{\ensuremath{\blacksquare}}\xspace}(x)}{1-a^-\xspace(x)}.
\end{equation}
Deleting the non-root outer edge of a rooted 2-dissection gives a
rooted quadrangulation (possibly degenerated). Taking into account the
marked inner vertex or edge, respectively chosen among the $n$ inner
vertices and $2n-1$ inner edges, we get:
\begin{equation}\label{eq:a}
\begin{cases}
a^{\scalebox{.4}{\ensuremath{\blacksquare}}\xspace}(x) ~=~ 2x + \sum_{n\geqslant 2} n q_n x^n ~=~ 2x + xq'(x) &\\
a^-\xspace(x) ~=~ 2x + \sum_{n\geqslant 2} (2n-1) q_n x^n ~=~ 2x + 2xq'(x) - q(x)&
\end{cases}
\end{equation}
Putting together Equations~\eqref{eq:d} and~\eqref{eq:a},
Corollary~\ref{cor} implies:
\begin{proposition}\label{prop:compquad}
The series $q(x)$ satisfies the following equation:
\begin{equation}\label{eq:q}
x \cdot [2q'(x)^2+3q'(x)+2] = q'(x) \cdot [1+q(x)].
\end{equation}
\end{proposition}
From this equation, written as $q'=x(2+2q'^2+3q')/(1+q)$,
one readily extracts the development of $q(x)$
incrementally:
\[
q(x)=x^2+2x^3+6x^4+22x^5+91x^6+408x^7+1938x^8+\dots
\]
As shown next, the exact expression of the coefficients (first obtained
by Tutte from the recursive method~\cite{Tu63} and subsequently
recovered by Schaeffer~\cite{S-these} using a bijection with ternary
trees) can also be recovered from~\eqref{eq:q}:
\begin{corollary}
For $n\geqslant 1$, the number of rooted simple quadrangulations with $n$
faces is equal to:
\[
\frac{4(3n)!}{n!(2n+2)!}.
\]
Equivalently, the series $q(x)$ is expressed as $q(x) = x
[\alpha(x)-2] [1-\alpha(x)]$, where $\alpha\equiv\alpha(x)$ is the
series of rooted ternary trees, specified by $\alpha = 1+x\alpha^3$.
\end{corollary}
\begin{proof}
Equation~\eqref{eq:q} above admits a unique power series solution
that is equal to $0$ at $0$, hence it suffices to check that $f
\equiv f(x) := x [\alpha(x)-2] [1-\alpha(x)]$ is indeed solution
of~\eqref{eq:q}; note that $x$ and $f(x)$ have rational expressions
in terms of $\alpha$, hence this also holds for $f'(x)$, since
$f'(x)=\frac{\mathrm{d}f}{\mathrm{d}\alpha} /
\frac{\mathrm{d}x}{\mathrm{d}\alpha}$. We have:
\[
x=\frac{\alpha-1}{\alpha^3},\qquad
f(x)=\frac{(\alpha-1)^2 (2-\alpha)}{\alpha^3},\quad
\text{and}\quad f'(x)=2\alpha-2.
\]
Plugging these expressions in Equation~\eqref{eq:q}, the two
handsides coincide, which concludes the proof.
As Alin Bostan showed us, Equation~\eqref{eq:q} can however be
solved directly, without guessing the solution. Let $r \equiv r(x) =
q'(x)$, Equation~\eqref{eq:q} rewrites:
\[
1 + q = x \cdot \frac{2r^2+3r+2}{r},
\]
hence, taking the derivative:
\begin{equation}\label{eq:deriv}
r = \frac{2r^2+3r+2}{r} + 2 x r' \frac{r^2-1}{r^2}, \qquad
\text{\emph{i.e.}\xspace} \qquad
(r+1) \cdot \left[ r(r+2) + 2 x r'(r-1) \right] = 0,
\end{equation}
or, assuming $r\neq 2$:
\begin{equation}\label{eq:alin}
(r+1) \cdot \frac{r^2}{(r+2)^2} \cdot \frac{\mathrm{d}}{\mathrm{d}x}
\left(\frac{x(r+2)^3}{r}\right) = 0 .
\end{equation}
Assuming now that $r\neq0$ and $r\neq 1$, we get $x (r(x)+2)^3 -
c r(x) = 0 $ for some constant $c$.
On the other hand, Equation~\eqref{eq:deriv} implies that
\[
\frac{\mathrm{d}}{\mathrm{d}x}\left(4q(x) - 2xr(x) + xr(x)^2\right)
= 0,
\qquad \text{hence} \qquad q(x) = \frac12 xr(x) + \frac14 xr(x)^2
+ c'
\]
for some constant~$c'$; $q(0) = r(0) = 0$ implies that $c'= 0$. Thus
the non affine solutions of Equation~\eqref{eq:q} have the following form:
\[
c + x \cdot (\alpha_c(x)-1)(2-\alpha_c(x)) \quad\text{where}\quad x
\alpha_c(x)^3 = c(\alpha_c(x)-1).
\]
Initial conditions for $q$ imply that $c=1$.
\end{proof}
\subsection{Getting a functional equation for simple triangulations}\label{sub:triangulation}
The case of simple triangulations is very similar to the one described
above for quadrangulations; only the decomposition is
more complicated.
Proposition~\ref{prop:quoclastrig} and Theorem~\ref{thm:bijtrig}
describe bijections between symmetric simple triangulations and two
families that are hence also in one-to-one correspondence:
\begin{corollary}\label{cor:trig}
Simple triangulations with $2n$ faces and one marked edge are
in one-to-one correspondence with quasi-simple pointed 1-dissections
with $2n$ faces.
\end{corollary}
Let $\ensuremath{\mathcal{T}}\xspace$ be the family of rooted simple triangulations and let
$t(x)=\sum_{n\geqslant 1}t_nx^n$ be the series of $\ensuremath{\mathcal{T}}\xspace$ according to half
the number of faces, \emph{i.e.}\xspace $t_n$ is the number of rooted simple
triangulations with $2n$ faces. We want to use
Corollary~\ref{cor:trig} to derive a functional equation for $t(x)$,
which requires to first express the family \ensuremath{\mathcal{S}}\xspace of simple
triangulations with a marked edge and the family $\ensuremath{\mathcal{D}}\xspace_3$ of quasi-simple
pointed 1-dissections in terms of~\ensuremath{\mathcal{T}}\xspace.
According to the Schnyder tree decomposition of simple triangulations,
each edge is canonically associated to one of the three outer
edges, hence simple triangulations with a marked edge are exactly
rooted simple triangulations with a marked edge chosen among the $n$
edges in the tree rooted at the root edge. Hence:
\begin{equation}\label{eq:trigleft}
s(x) = \sum_{n\geqslant 1} n t_n x^{n} = xt'(x).
\end{equation}
The enumeration of quasi-simple pointed triangulations can be carried
out similarly to quasi-simple pointed quadrangulations, but
extra care has to be taken in order to deal with special conditions
for both loops and 2-cycles. Observe that any 1-dissection \scalebox{.8}{\ensuremath{D}}\xspace in
$\ensuremath{\mathcal{D}}\xspace_3$ is implicitly rooted, and that removing the outer loop yields
either a unique edge or a quasi-simple pointed triangular 2-dissection that
can be canonically rooted with the same root vertex as \scalebox{.8}{\ensuremath{D}}\xspace,
with the additional constraint that there is no loop
incident to the root vertex; let us call $\ensuremath{\mathcal{U}}\xspace$ this family and $\ensuremath{u}\xspace$
its generating series according to half the number of triangular
faces. Then the generating series $d_3$ of $\ensuremath{\mathcal{D}}\xspace_3$ is given by:
\[
d_3(x) = x \cdot [1+\ensuremath{u}\xspace(x)].
\]
Note that removing the non-root edge incident to the outer face of a
triangular 2-dissection produces a rooted triangulation with one face
less. Hence families of rooted triangular 2-dissections (such as \ensuremath{\mathcal{U}}\xspace)
and families of rooted triangulations can be naturally identified. To
decompose elements of \ensuremath{\mathcal{U}}\xspace, let \ensuremath{\mathcal{V}}\xspace be the family of quasi-simple
triangular 2-dissections and \ensuremath{\mathcal{T}^\carre}\xspace, \ensuremath{\mathcal{T}^\trait}\xspace and \ensuremath{\mathcal{T}^{\bullet\!\!\trait}}\xspace be respectively the
families of rooted simple triangulations (or equivalently, triangular
2-dissections with neither loop nor 2-cycle except the outer
face) with respectively a marked inner vertex, a marked inner edge,
and a marked inner edge incident to the root vertex. Let us also
denote \ensuremath{v}\xspace, \ensuremath{t^\carre}\xspace, \ensuremath{t^\trait}\xspace and \ensuremath{t^{\bullet\!\!\trait}}\xspace the corresponding generating series
according to half the number of triangular faces.
\medskip
\begin{figure}
\centering
\subfigure[]{\includegraphics[scale=.9,page=1]{Trig_U}}\qquad
\subfigure[]{\includegraphics[scale=.9,page=2]{Trig_U}}\qquad
\subfigure[]{\includegraphics[scale=.9,page=3]{Trig_U}}\qquad
\subfigure[]{\includegraphics[scale=.9,page=4]{Trig_U}}
\caption{The four cases in the decomposition of an element of \ensuremath{\mathcal{U}}\xspace.}
\label{fig:Trig_U}
\end{figure}
Let us now proceed to the decomposition of an element of $\ensuremath{\mathcal{U}}\xspace$, four
different configurations can appear (see Figure~\ref{fig:Trig_U}):
\begin{itemize}
\item[(a)] either the outer face is the only 2-cycle;
\item[(b)] or the outer face is incident to a single face, the third
side of which is a loop;
\item[(c)] or the outermost 2-cycle (other than the outer face) is
incident to the root vertex;
\item[(d)] or the outermost 2-cycle (other than the outer face) is not
incident to the root vertex.
\end{itemize}
Note that each inner edge of a simple triangulation, and in particular
of any map \scalebox{.8}{\ensuremath{T}}\xspace in $\ensuremath{\mathcal{T}^{\bullet\!\!\trait}}\xspace$ is naturally oriented according to the
minimal 3-orientation of~\scalebox{.8}{\ensuremath{T}}\xspace.
Hence there is a canonical way of plugging a rooted 2-dissection in
the (rooted) 2-cycle obtained by \emph{opening} the marked edge of a
map in \ensuremath{\mathcal{T}^\trait}\xspace. Hence:
\[ \ensuremath{\mathcal{U}}\xspace = \ensuremath{\mathcal{T}^\carre}\xspace + \ensuremath{\mathcal{X}}\xspace \cdot (1+\ensuremath{\mathcal{U}}\xspace) + \ensuremath{\mathcal{T}^{\bullet\!\!\trait}}\xspace \cdot \ensuremath{\mathcal{U}}\xspace + (\ensuremath{\mathcal{T}^\trait}\xspace \setminus
\ensuremath{\mathcal{T}^{\bullet\!\!\trait}}\xspace)\cdot \ensuremath{\mathcal{V}}\xspace,
\]
where \ensuremath{\mathcal{X}}\xspace is an atom representing a triangle. A similar decomposition
can be written for \ensuremath{\mathcal{V}}\xspace, and putting things together and translating
them into the language of generating series, we obtain:
\[
\begin{cases}
\ensuremath{u}\xspace = \ensuremath{t^\carre}\xspace + x(1+\ensuremath{u}\xspace) + \ensuremath{t^{\bullet\!\!\trait}}\xspace\ensuremath{u}\xspace +(\ensuremath{t^\trait}\xspace - \ensuremath{t^{\bullet\!\!\trait}}\xspace)\ensuremath{v}\xspace,&\\
\ensuremath{v}\xspace = \ensuremath{t^\carre}\xspace + 2x(1+\ensuremath{u}\xspace) + \ensuremath{t^\trait}\xspace\ensuremath{v}\xspace.&
\end{cases}
\]
The series $\ensuremath{t^\carre}\xspace$, $\ensuremath{t^\trait}\xspace$ and $\ensuremath{t^{\bullet\!\!\trait}}\xspace$can now be expressed in
terms of $t$ as follows. A triangulation with $2n$ faces has $3n$
edges and $n+2$ vertices, so removing the non-root outer edge and
taking into account the marked inner vertex or edge, we get:
\[
\begin{cases}
\ensuremath{t^\carre}\xspace(x)=\sum_{n\geqslant 1}nt_nx^n=xt'(x), &\\
\ensuremath{t^\trait}\xspace(x)=\sum_{n\geqslant 1}(3n-1)t_nx^n=3xt'(x)-t(x).&
\end{cases}
\]
The computation of $\ensuremath{t^{\bullet\!\!\trait}}\xspace(x)$ is a little more complicated. Let \scalebox{.8}{\ensuremath{T}}\xspace be
a simple rooted triangulation with at least four faces. By merging the
two ends of the root edge of \scalebox{.8}{\ensuremath{T}}\xspace and collapsing the inner triangle
incident to it into an edge, we obtain a 2-dissection that is
canonically rooted with same root vertex as \scalebox{.8}{\ensuremath{T}}\xspace, and with a marked
edge incident to that vertex. This triangulation is not far from being
simple: it has no loop and the only 2-cycles separate the two marked
edges. Hence, decomposing this map along the sequence of nested
cycles, we get a sequence of elements of~\ensuremath{\mathcal{T}^{\bullet\!\!\trait}}\xspace, hence:
\[
\frac{t(x)-x}{x}= \sum_{n\geqslant 2}t_nx^n = \frac{\ensuremath{t^{\bullet\!\!\trait}}\xspace(x)}{1-\ensuremath{t^{\bullet\!\!\trait}}\xspace(x)},\quad
\text{from which we obtain}\quad \ensuremath{t^{\bullet\!\!\trait}}\xspace(x)=\frac{t(x)-x}{t(x)}.
\]
Combining all these equations and with the help of computer algebra,
we obtain the following expression for the generating series of rooted
quasi-simple triangular 1-dissections:
\begin{equation}\label{eq:trigright}
d_3(x) = \frac{x \cdot [1+t(x)-2xt'(x)]}{1-2x+2t(x)-3xt'(x)+t(x)^2-3xt'(x)t(x)}.
\end{equation}
The following result is then obtained by gathering
Corollary~\ref{cor:trig} and Equations~\eqref{eq:trigleft}
and~\eqref{eq:trigright}, after some easy simplifications:
\begin{proposition}
The generating series $t(x)$ of rooted simple triangulations
according to half the number of faces is the unique solution of the
following equation:
\begin{equation}\label{eq:t}
3x t'(x)^2 + 1 = [t(x) + 1] \cdot t'(x),
\end{equation}
which takes value 0 at 0.
\end{proposition}
From this equation, written as $t'=(3xt'^2+1)/(t+1)$, one readily
extracts the development of $t(x)$ incrementally:
\[
t(x)=x+x^2+3x^3+13x^4+68x^5+399x^6+2530x^7+16965x^8+\dots
\]
As in the quadrangular case, the exact expression of the coefficients
(first obtained by Tutte from the recursive method~\cite{Tu63} and
subsequently recovered by Poulalhon and Schaeffer~\cite{S-these} using
a bijection with some decorated trees that are in bijection with
quaternary trees) can also be recovered from~\eqref{eq:t}:
\begin{corollary}
For $n\geqslant 1$, the number of rooted simple triangulations with $2n$
faces is equal to:
\[
\frac{2(4n-3)!}{n!(3n-1)!}.
\]
Equivalently, the series $t(x)$ is expressed as $t(x)=
\dfrac{(\alpha(x)-2)(1-\alpha(x))}{\alpha(x)^2}$, where
$\alpha\equiv\alpha(x)$ is the series of rooted quaternary trees,
specified by $\alpha=1+x\alpha^4$.
\end{corollary}
\begin{proof}
Equation~\eqref{eq:t} above admits a unique power series solution
that is equal to $0$ at $0$, so it suffices to check that $f\equiv
f(x):=(\alpha(x)-2)(1-\alpha(x)) / \alpha(x)^2$ is solution of~\eqref{eq:t}.
Note that $x$ and $f(x)$ have rational expressions in terms of
$\alpha$, and so does $f'(x)=\frac{\mathrm{d}f}{\mathrm{d}\alpha} /
\frac{\mathrm{d}x}{\mathrm{d}\alpha}$:
\[
x=\frac{\alpha-1}{\alpha^4},\qquad f(x) =
\dfrac{(\alpha(x)-2)(1-\alpha(x))}{\alpha(x)^2},\quad
\text{and}\quad f'(x)=\alpha^2,
\]
and these expressions satisfy Equation~\eqref{eq:t}, which concludes
the proof.
Now as for Equation~\eqref{eq:q}, a direct proof without guessing
the solution is possible; first let $r(x) = t'(x)$,
Equation~\eqref{eq:t} rewrites $1+t = 3xr + 1/r$, hence $r = 3r'+3r
-r'/r^2$, \emph{i.e.}\xspace $u'(3xu^2-1) + 2u^3 = 0$. Then we seek for $A$, $B$
such that $\frac{\mathrm d}{\mathrm d x} \left(A(r) x + B(r) \right)
= 0$. Easy computations show that $x r^{3/2} + r^{-1/2} $ is
suitable. Using initial conditions, we get $x r^2+1 = r^{1/2}$, \emph{i.e.}\xspace,
with $\alpha = x r^2 + 1$, $1+x\alpha^4 = \alpha$.
\end{proof}
\section{Radial distance of symmetric quadrangular dissections}\label{sec:quad_dista}
\label{sec:dist}
For $k\geqslant 2$, $i>0$, and \ensuremath{\cD^{(k)}}\xspace a family of $k$-symmetric dissections,
let \ensuremath{\cD_i^{(k)}}\xspace be the family of dissections in \ensuremath{\cD^{(k)}}\xspace where the central
vertex is at distance $i$ from the outer face boundary; define the
size of a $k$-symmetric quadrangular (\emph{resp.}\xspace triangular) dissection \scalebox{.8}{\ensuremath{D}}\xspace
as the integer $n$ such that \scalebox{.8}{\ensuremath{D}}\xspace has $kn$ inner faces (\emph{resp.}\xspace $(2n+1)k$
inner faces). Let $\ensuremath{D^{(k)}_i}\xspace(x)$ be the generating series of \ensuremath{\cD_i^{(k)}}\xspace with
respect to the size.
We compute here the expression of $\ensuremath{D^{(k)}_i}\xspace(x)$ for general, simple, and
irreducible $k$-symmetric quadrangular $2k$-dissections and triangular
$k$-dissections. So from now on \ensuremath{\cD^{(k)}}\xspace is either a family of
$k$-symmetric quadrangular $2k$-dissections or a family of
$k$-symmetric triangular $k$-dissections. Quadrangular and triangular
dissections are treated respectively in this section and in the next
one (Section~\ref{sec:dist_trig}). Note that, when $k=2$, $\ensuremath{D^{(k)}_i}\xspace(x)=0$
for quadrangular irreducible dissections and for triangular simple and
irreducible dissections, and when $k=3$, $\ensuremath{D^{(k)}_i}\xspace(x)=0$ for triangular
irreducible dissections.
We use the letter $\ensuremath{\mathcal{F}}\xspace$, $\ensuremath{\mathcal{G}}\xspace$, $\ensuremath{\mathcal{H}}\xspace$ (instead of $\ensuremath{\mathcal{D}}\xspace$) for the
general, simple, and irreducible case respectively. To obtain the
generating function expressions, we combine results of Bouttier
\emph{et al.}\xspace~\cite{BoDFGu03,BoGu} on the $2$-point functions of general
quadrangulations
with quotient and substitution operations, and
Lemma~\ref{lem:quotient_distance}.
\begin{rem}\label{rk:encloses}
Lemma~\ref{lem:quotient_distance} implies that a $k$-symmetric
quadrangular dissection has no cycle of length less than $2k$ that
strictly encloses the central vertex (indeed the $k$-quotient has
only faces of even degree, hence is bipartite, hence has no loop).
And a $k$-symmetric triangular dissection has no cycle of length
less than $k$ that strictly encloses the central vertex.
\end{rem}
\subsection{Symmetric quadrangular dissections}
Define the algebraic generating function $P\equiv P(x)$ by
\begin{equation}\label{eq:defP}
P=1+3xP^2,
\end{equation}
and define the algebraic generating function $X\equiv X(x)$ by
\begin{equation}\label{eq:defX}
X+\frac{1}{X}+1=\frac{3}{P-1}.
\end{equation}
Define also $X_{\infty}:=P$. The following result is very closely
related to a result by Bouttier, Di Francesco, and Guitter~\cite{BoDFGu03}.
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 2$, the generating function $\ensuremath{F^{(k)}_i}\xspace(x)$
has the following expression (which does not depend on $k$):
\[
\ensuremath{F^{(k)}_i}\xspace(x)=X_{i+1}-X_i,\ \ \ \mathrm{where}\ X_i=X_{\infty}\frac{(1-X^i)(1-X^{i+3})}{(1-X^{i+1})(1-X^{i+2})}.
\]
\end{proposition}
\begin{proof}
If we take the $k$-quotient of dissection in $\ensuremath{\mathcal{F}}\xspace_i^{(k)}$ we obtain
a quadrangular 2-dissection with a pointed inner vertex
at distance $i$ from the outer $2$-gon, according to
Lemma~\ref{lem:quotient_distance}. Notice that the outer $2$-cycle
can be contracted (on the sphere) into a single edge $e$. This yields a
quadrangulation of the sphere with a marked edge $e$ and pointed
vertex $v$ at distance $i$ from $e$ (i.e., the extremity of $e$
closest from $v$ is at distance $i$ from $v$). Hence $\ensuremath{F^{(k)}_i}\xspace(x)$ is equal to the
generating function $F_i(x)$ of quadrangulations with a marked
edge $e$ and a marked vertex $v$ at distance $i$ from $e$. The algebraic
expression of $\ensuremath{F^{(k)}_i}\xspace(x)=F_i(x)$, as $X_{i+1}-X_i$, has been obtained
by Bouttier \emph{et al.}\xspace~\cite{BoDFGu03}.
\end{proof}
\subsection{Symmetric simple quadrangular dissections}
\begin{figure}
\centering
\subfigure[The maximal 2-cycles,]{\includegraphics[page=1]{simpleCore}}
\qquad\qquad
\subfigure[and the simple core.]{\includegraphics[page=2]{simpleCore}}
\caption{A 3-symmetric quadrangulation and its simple 3-symmetric core, obtained
by collapsing all maximal 2-cycles into edges.}
\label{fig:simplecore}
\end{figure}
We now compute $y\mapsto\ensuremath{G^{(k)}_i}\xspace(y)$. We classically proceed by
substitution (a substitution approach
is also discussed in~\cite{BoGu} for rooted quadrangulations).
The \emph{core} of $\gamma\in\ensuremath{\cF_i^{(k)}}\xspace$ is
obtained by collapsing each maximal $2$-cycle (for the enclosed area)
of $\gamma$ into a single edge (note that no $2$-cycle can strictly
enclose the central vertex, according to Remark~\ref{rk:encloses}),
see Figure~\ref{fig:simplecore}. This yields a simple dissection in
$\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$, and the distance of the central vertex to the outer boundary
is still $i$ (because there is no way of shortening this distance by
travelling inside a $2$-cycle). Conversely each $\gamma\in\ensuremath{\cF_i^{(k)}}\xspace$ is
uniquely obtained from $\kappa\in\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$ ---with $nk$ inner faces---
where each of the $(2n+1)k$ edges is either left unchanged or blown
into a double edge in the interior of which a rooted quadrangulation
is patched, in a way that respects the symmetry of order $k$ (that is,
the $k$ edges in an orbit of edges of $\kappa$ undergo the same
substitution operation). Denoting by $f\equiv f(x)$ the series of
rooted quadrangulations according to the number of faces, we obtain
$\ensuremath{F^{(k)}_i}\xspace(x)=\sum_{n\geqslant 1}[y^n]\ensuremath{G^{(k)}_i}\xspace(y)\cdot x^n\cdot(1+f)^{2n+1}$. In other
words,
\begin{equation}\label{eq:FGquad}
\ensuremath{F^{(k)}_i}\xspace(x)=(1+f)\cdot\ensuremath{G^{(k)}_i}\xspace(x\cdot(1+f)^2).
\end{equation}
The series $f=f(x)$ is well known to be algebraic,
having a rational expression in terms of $P$: $f=P(4-P)/3-1$ (we will
need this expression a few times).
Define $Q\equiv Q(y)$ as the algebraic series in $y$ defined by
\begin{equation}
Q=1+yQ^3.
\end{equation}
\begin{lemma}\label{lem:xy_quad}
Let $K(x)$ and let $L(y)$ be related by $K(x)=L(x(1+f)^2)$. Let
$\widehat{K}(P)$ and $\widehat{L}(Q)$ be the expressions of $K(x)$
and $L(y)$ in terms of $P\equiv P(x)$ and $Q\equiv Q(y)$, i.e.,
$K(x)=\widehat{K}(P(x))$ and $L(y)=\widehat{L}(Q(y))$. Then
\[
\widehat{L}(Q)=\widehat{K}\Big(4-3/Q\Big).
\]
\end{lemma}
\begin{proof}
The change of variable relation is $y=x(1+f)^2$. We have
\[
y=x(1+f)^2=\frac{P-1}{3P^2}\Big(P\frac{4-P}{3}\Big)^2=\frac{P-1}{3}\Big(\frac{4-P}{3}\Big)^2
\]
Hence, if we write $Q\equiv Q(y)=3/(4-P(x))$, we have
$y=(1-1/Q)Q^{-2}$, so that $Q=1+yQ^3$. In addition $P(x)=4-3/Q(y)$.
\end{proof}
For the rest of this subsection, when we have $y$ and $x$ together in
an equation we assume $y$ and $x$ to be related by $y=x(1+f)^2$.
Define $Y_{\infty}(y):=X_{\infty}(x)/(1+f)$, $Y_i(y):=X_i(x)/(1+f)$,
and $Y(y):=X(x)$. The expression of $X_i(x)$ in terms of
$X_{\infty}(x)$ and $X(x)$ ensures that (we do not need
Lemma~\ref{lem:xy_quad} at this step):
\[
Y_i=Y_{\infty}\frac{(1-Y^i)(1-Y^{i+3})}{(1-Y^{i+1})(1-Y^{i+2})}.
\]
In addition, since $\ensuremath{G^{(k)}_i}\xspace(y)=\ensuremath{F^{(k)}_i}\xspace(x)/(1+f)$, we have
\[
\ensuremath{G^{(k)}_i}\xspace(y)=Y_{i+1}-Y_i.
\]
We now apply Lemma~\ref{lem:xy_quad} to get an algebraic expression of
$\ensuremath{G^{(k)}_i}\xspace(y)$, written in terms of $Q(y)$. We have
\[
Y_{\infty}(y)=\frac{X_{\infty}(x)}{1+f(x)}=\frac{3}{4-P}.
\]
Then Lemma~\ref{lem:xy_quad} ensures that $Y_{\infty}(y)=Q(y)$.
Lemma~\ref{lem:xy_quad} and the relation $X+1/X+1=3/(P-1)$ ensure
that $Y\equiv Y(y)$ is the algebraic generating function specified by
\begin{equation}\label{eq:Yquad}
Y+\frac{1}{Y}=\frac{1}{Q-1}.
\end{equation}
To summarize we obtain:
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 2$, the generating function $\ensuremath{G^{(k)}_i}\xspace(y)$
has the expression (with $Y_{\infty}=Q$):
\[
\ensuremath{G^{(k)}_i}\xspace(y)=Y_{i+1}-Y_i,\ \ \ \mathrm{where}\ Y_i=Y_{\infty}\frac{(1-Y^i)(1-Y^{i+3})}{(1-Y^{i+1})(1-Y^{i+2})}.
\]
\end{proposition}
\subsection{Symmetric irreducible quadrangular dissections}
\begin{figure}
\centering
\subfigure[The maximal non empty 4-cycles,]{\qquad\includegraphics[page=1]{irreductibleCore}\qquad}
\qquad
\subfigure[and the irreductible core.]{\includegraphics[page=2]{irreductibleCore}}\qquad
\caption{A 3-symmetric simple quadrangulation and its irreductible 3-symmetric core, obtained
by emptying all maximal 4-cycles into faces.}
\label{fig:irrcore}
\end{figure}
We now use a substitution approach at faces (instead of edges) to get
an expression for $\ensuremath{H^{(k)}_i}\xspace(z)$ for $k\geqslant 3$ and $i>0$.
The \emph{core} of $\gamma\in\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$ is obtained by emptying each
maximal (for the enclosed area) $4$-cycle of $\gamma$ (note that no
$4$-cycle strictly encloses the central vertex, according to
Remark~\ref{rk:encloses}), see Figure~\ref{fig:irrcore}. This yields a symmetric irreducible
dissection $\kappa\in\ensuremath{\ensuremath{\mathcal{H}}\xspace_i^{(k)}}\xspace$, and the distance of the pointed vertex to
the outer boundary is still $i$ (because there is no way of shortening
this distance by travelling inside a $4$-cycle). Conversely each
$\gamma\in\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$ is uniquely obtained from $\kappa\in\ensuremath{\ensuremath{\mathcal{H}}\xspace_i^{(k)}}\xspace$
where at each face a rooted simple quadrangulation
(with at least one inner face) is patched, in a way that respects the
symmetry of order $k$ (i.e., the $k$ faces of an orbit undergo the
same patching operation). Denoting by $g\equiv g(y)$ the series of
rooted simple non-degenerated quadrangulations according to the number of inner faces, we
obtain, for $k\geqslant 3$ and $i>0$:
\[
\ensuremath{G^{(k)}_i}\xspace(y)=\ensuremath{H^{(k)}_i}\xspace(g(y)).
\]
Again the series $g\equiv g(y)$ is well known to be algebraic, having a rational
expression in terms of $Q$:
$g=3Q-Q^2-2$.
Define the algebraic series $R\equiv R(z)$ by
\begin{equation}
R=z+R^2.
\end{equation}
\begin{lemma}\label{lem:yz_quad}
Let $L(y)$ and let $M(z)$ be related by $L(y)=M(g(y))$. Let
$\widehat{L}(Q)$ and $\widehat{M}(R)$ be the expressions of $L(y)$
and $M(z)$ in terms of $Q\equiv Q(y)$ and $R\equiv R(z)$, i.e.,
$L(y)=\widehat{L}(Q(y))$ and $M(z)=\widehat{M}(R(z))$. Then
\[
\widehat{M}(R)=\widehat{L}(R+1).
\]
\end{lemma}
\begin{proof}
The change of variable relation is $z=g(y)$. We have
\[
z=g(y)=3Q-Q^2-2=-(Q-1)^2+(Q-1)
\]
Hence, if we write $R\equiv R(z)=Q(y)-1$, we have $z=-R^2+R$, so
that $R=z+R^2$. In addition $Q(y)=R(z)+1$.
\end{proof}
For the rest of this subsection, when we have $z$ and $y$ together in
an equation we assume $z$ and $y$ to be related by $z=g(y)$. Define
$Z_{\infty}(z):=Y_{\infty}(y)$, $Z_i(z):=Y_i(y)$, and $Z(z):=Y(y)$.
The expression of $Y_i(y)$ in terms of $Y_{\infty}(y)$ and $Y(y)$
ensures that (we do not need Lemma~\ref{lem:yz_quad} at this step):
\[
Z_i=Z_{\infty}\frac{(1-Z^i)(1-Z^{i+3})}{(1-Z^{i+1})(1-Z^{i+2})}.
\]
In addition, since $\ensuremath{H^{(k)}_i}\xspace(z)=\ensuremath{G^{(k)}_i}\xspace(y)$, we have
\[
\ensuremath{H^{(k)}_i}\xspace(z)=Z_{i+1}-Z_i.
\]
We now apply Lemma~\ref{lem:yz_quad} to get an algebraic expression of
$\ensuremath{H^{(k)}_i}\xspace(z)$, written in terms of $R(z)$. We have
\[
Z_{\infty}(z)=Y_{\infty}(y)=Q(y)=R(z)+1.
\]
Lemma~\ref{lem:yz_quad} and the relation $Y+1/Y=1/(Q-1)$ also ensure
that $Z\equiv Z(z)$ is the algebraic generating function specified by
\begin{equation}\label{eq:Zquad}
Z+\frac{1}{Z}=\frac{1}{R}.
\end{equation}
To summarize we obtain:
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 3$, the generating function $\ensuremath{H^{(k)}_i}\xspace(z)$
has the expression (with $Z_{\infty}=R+1$):
\[
\ensuremath{H^{(k)}_i}\xspace(z)=Z_{i+1}-Z_i,\ \ \ \mathrm{where}\ Z_i=Z_{\infty}\frac{(1-Z^i)(1-Z^{i+3})}{(1-Z^{i+1})(1-Z^{i+2})}.
\]
\end{proposition}
\section{Radial distance of symmetric triangular dissections}\label{sec:dist_trig}
\subsection{Symmetric triangular dissections}
Define the algebraic generating function $P\equiv P(x)$ by
\begin{equation}\label{eq:defP_triang}
P^2=1+8xP^3,
\end{equation}
and define the algebraic generating function $X\equiv X(x)$ by
\begin{equation}\label{eq:defX_triang}
X+\frac{1}{X}+2=\frac{8}{P^2-1}.
\end{equation}
Define also $X_{\infty}\equiv X_{\infty}(x)$ and $A_{\infty}\equiv A_{\infty}(x)$ by
\begin{equation}
X_{\infty}=P,\ \ A_{\infty}\ \!\!^2=\frac{2P(P-1)}{(1+P)}.
\end{equation}
As for quadrangulations the following result is very closely
related to a result by Bouttier and Guitter~\cite{BoGu12}.
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 2$, the generating function $\ensuremath{F^{(k)}_i}\xspace(x)$ has the following
expression (which does not depend on $k$):
\[
\ensuremath{F^{(k)}_i}\xspace(x)=X_{i+1}-X_{i-1}+A_i\ \!\!^2-A_{i-1}\ \!\!^2,
\]
where
\[
X_i = X_{\infty} \cdot \frac{(1-X^i)(1-X^{i+2})}{(1-X^{i+1})^2}, \quad
A_i=A_{\infty}\cdot\Big(1-\frac{P+1}{4}X^i\frac{(1-X)(1-X^2)}{(1-X^{i+1})(1-X^{i+2})}\Big).
\]
\end{proposition}
\begin{proof}
We recall the result in~\cite{BoGu12} about triangulations:
\begin{itemize}
\item the series $U_i$ of triangulations of the sphere with a
pointed vertex $u$ and a marked edge $e=(v,v')$ with $v$ at
distance $i$ and $v'$ at distance $i-1$ from $u$ is given by
$U_i=X_i-X_{i-1}$,
\item the series $V_i$ of triangulations of the sphere with a
pointed vertex $u$ and a marked oriented edge $e=(v,v')$ with $v$
and $v'$ at distance $i$ from $u$ is given by $V_i=A_i\
\!\!^2-A_{i-1}\ \!\!^2$. (Note that the series $V_i$ decomposes as
$V_{i,\mathrm{loop}}+V_{i,\mathrm{distinct}}$, whether the
extremities of the marked edge are equal or distinct.)
\end{itemize}
The $k$-quotient of a dissection in $\ensuremath{\mathcal{F}}\xspace_i^{(k)}$ is a triangular
1-dissection $D$ with a pointed inner vertex $u$ at
distance $i$ from the outer loop, according to
Lemma~\ref{lem:quotient_distance}. The vertex incident to the outer
loop is called the \emph{root-vertex} and denoted by $v$. If we
delete the outer loop we obtain a map $\widetilde{D}$ with an outer
face of degree $2$ and all inner faces of degree $3$. Two cases can
occur: either the outer face contour of $\widetilde{D}$ is a $2$-gon
or is made of two adjacent loops. In the first case let $v'$ be the
other vertex of the $2$-gon: $v'$ is at distance either $i-1$,
$i+1$, or $i$ from $u$, giving respective contributions
$U_i$, $U_{i+1}$, $V_{i,\mathrm{distinct}}$. In the second case we have an
ordered pair $D_1,D_2$ of 1-dissections, and one of
these two dissections contains a marked vertex at distance $i$ from
$v$. Orient the outer loop of $D_1$ clockwise and the outer loop of
$D_2$ counterclockwise, and paste $D_1$ and $D_2$ together at
their outer loops. This yields a triangulation of the sphere with a
pointed vertex $u$ and a marked oriented loop whose incident vertex
is at distance $i$ from $u$. The corresponding contribution is
$V_{i,\mathrm{loop}}$. Gathering all cases we obtain
$\ensuremath{F^{(k)}_i}\xspace=U_i+U_{i+1}+V_{i,\mathrm{distinct}}+V_{i,\mathrm{loop}}=U_i+U_{i+1}+V_i$.
\end{proof}
\subsection{Symmetric simple triangular dissections}
Call a rooted triangulation \emph{simply-rooted} if the root-edge is
not a loop. To compute $y\mapsto\ensuremath{G^{(k)}_i}\xspace(y)$, we proceed very similarly
as for quadrangulations. Each $\gamma\in\ensuremath{\cF_i^{(k)}}\xspace$ (for $k\geqslant 3$) is
uniquely obtained from $\kappa\in\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$ ---with $(2n+1)k$ inner
faces--- where each of the $(3n+2)k$ edges is either left unchanged or
blown into a double edge in the interior of which a simply-rooted
triangulation is patched, in a way that respects the symmetry of order
$k$ (that is, the $k$ edges in an orbit of edges of $\kappa$ undergo
the same substitution operation). Denoting by $f\equiv f(x)$ the
series of simply-rooted triangulations according to half the number of
faces, we obtain $\ensuremath{F^{(k)}_i}\xspace(x)=\sum_{n\geqslant 1}[y^n]\ensuremath{G^{(k)}_i}\xspace(y)\cdot
x^n\cdot(1+f)^{3n+2}$. In other words,
\begin{equation}\label{eq:FGtriang}
\ensuremath{F^{(k)}_i}\xspace(x)=(1+f)^2\cdot\ensuremath{G^{(k)}_i}\xspace(x\cdot(1+f)^3).
\end{equation}
The series $f=f(x)$ is algebraic,
with a rational expression in terms of $P$: $1+f=-P(P^2-9)/8$.
Define $Q\equiv Q(y)$ as the algebraic series in $y$ given by
\begin{equation}
Q=\frac{y}{(1-Q)^3},
\end{equation}
and define also $\ensuremath{\widetilde{Q}}\xspace\equiv \ensuremath{\widetilde{Q}}\xspace(y)$ as $\ensuremath{\widetilde{Q}}\xspace:=(1+8Q)^{1/2}$.
\begin{lemma}\label{lem:xy_triang}
Let $K(x)$ and $L(y)$ be related by $K(x)=L(x(1+f)^3)$. Let
$\widehat{K}(P)$ and $\widetilde{L}(\ensuremath{\widetilde{Q}}\xspace)$ be the expressions of
$K(x)$ and $L(y)$ in terms of $P\equiv P(x)$ and $\ensuremath{\widetilde{Q}}\xspace\equiv \ensuremath{\widetilde{Q}}\xspace(y)$,
i.e., $K(x)=\widehat{K}(P(x))$ and
$L(y)=\widetilde{L}(\ensuremath{\widetilde{Q}}\xspace(y))$. Then the expressions are the same,
i.e.,
\[
\widetilde{L}=\widehat{K}.
\]
\end{lemma}
\begin{proof}
The change of variable relation is $y=x(1+f)^3$. We have:
\[
y=x(1+f)^3=-\frac{1}{2^{12}}(P^2-1)(P^2-9)^3.
\]
Hence, if we write $Q\equiv Q(y)=(P(x)^2-1)/8$,
we have $y=Q(1-Q)^3$. In addition $P(x)=\ensuremath{\widetilde{Q}}\xspace(y)$, hence $\widetilde{L}=\widehat{K}$.
\end{proof}
Define (with $y$ and $x$ related by $y=x(1+f)^3$):
\[
Y_{\infty}(y):=\frac{X_{\infty}(x)}{(1+f)^2},\ Y_i(y):=\frac{X_i(x)}{(1+f)^2},\ Y(y):=X(x),
\]
and define
\[
B_{\infty}(y)=\frac{A_{\infty}(x)}{(1+f)},\ B_i(y):=\frac{A_i(x)}{(1+f)}.
\]
The expression of $X_i(x)$ in terms of $X_{\infty}(x)$ and $X(x)$
ensures that
\[
Y_i=Y_{\infty}\frac{(1-Y^i)(1-Y^{i+2})}{(1-Y^{i+1})^2}.
\]
The expression of $A_i(x)$ in terms of $A_{\infty}(y)$ and $X(x)$ and
Lemma~\ref{lem:xy_triang} (to replace $P$ by $\ensuremath{\widetilde{Q}}\xspace$ in the expression)
ensure that
\[
B_i=B_{\infty}\cdot\Big(1-\frac{\ensuremath{\widetilde{Q}}\xspace+1}{4}Y^i\frac{(1-Y)(1-Y^2)}{(1-Y^{i+1})(1-Y^{i+2})}\Big).
\]
In addition, since $\ensuremath{G^{(k)}_i}\xspace(y)=\ensuremath{F^{(k)}_i}\xspace(x)/(1+f)^2$, we have
\[
\ensuremath{G^{(k)}_i}\xspace(y)=Y_{i+1}-Y_{i-1}+B_i\ \!\!^2-B_{i-1}\ \!\!^2.
\]
Using Lemma~\ref{lem:xy_triang} we obtain (after simplifications)
\begin{equation}
Y_{\infty}=\frac{1}{\ensuremath{\widetilde{Q}}\xspace(1-Q)^2},\ \ B_{\infty}\ \!\!^2=\frac{16\cdot Q}{\ensuremath{\widetilde{Q}}\xspace(1+\ensuremath{\widetilde{Q}}\xspace)^2(1-Q)^2}.
\end{equation}
Lemma~\ref{lem:xy_triang} and the relation $X+1/X+1=8/(P^2-1)$ also
ensure that $Y\equiv Y(y)$ is the algebraic generating function
specified by
\begin{equation}\label{eq:Ytriang}
Y+\frac{1}{Y}+2=\frac{1}{Q}.
\end{equation}
To summarize we obtain:
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 3$, the generating function $\ensuremath{G^{(k)}_i}\xspace(y)$
has the following expression (which does not depend on $k$):
\[
\ensuremath{G^{(k)}_i}\xspace(y)=Y_{i+1}-Y_{i-1}+B_i\ \!\!^2-B_{i-1}\ \!\!^2,
\]
where
\[
Y_i=Y_{\infty}\frac{(1-Y^i)(1-Y^{i+2})}{(1-Y^{i+1})^2},\ \
B_i=B_{\infty}\cdot\Big(1-\frac{\ensuremath{\widetilde{Q}}\xspace+1}{4}Y^i\frac{(1-Y)(1-Y^2)}{(1-Y^{i+1})(1-Y^{i+2})}\Big).
\]
\end{proposition}
\subsection{Symmetric irreducible triangular dissections}
We now use a substitution approach at faces to get an expression for
$\ensuremath{H^{(k)}_i}\xspace(z)$ for $k\geqslant 4$ and $i>0$. Similarly as in the quadrangulated
case, each $\gamma\in\ensuremath{\ensuremath{\mathcal{G}}\xspace_i^{(k)}}\xspace$ (for $k\geqslant 4$) is uniquely obtained from
$\kappa\in\ensuremath{\ensuremath{\mathcal{H}}\xspace_i^{(k)}}\xspace$ ---with $(2n+1)k$ inner faces--- where at each face a
rooted simple triangulation (with at least one inner face) is patched,
in a way that respects the symmetry of order $k$ (i.e., the $k$ faces
of an orbit undergo the same patching operation). Denoting by
$g\equiv g(y)$ the series of rooted simple triangulations according to
half the number of faces, we obtain, for $k\geqslant 4$ and $i>0$:
\[
\ensuremath{G^{(k)}_i}\xspace(y)=\frac{g}{y}\ensuremath{H^{(k)}_i}\xspace(g^2/y).
\]
Again the series $g\equiv g(y)$ is well known to be algebraic, having a rational
expression in terms of $Q$:
$g=Q-2Q^2$.
Define the algebraic series $R\equiv R(z)$ by
\begin{equation}
R=\frac{z}{(1-R)^2}.
\end{equation}
Define also $\ensuremath{\widetilde{R}}\xspace\equiv\ensuremath{\widetilde{R}}\xspace(z)$ as $\ensuremath{\widetilde{R}}\xspace:=\sqrt{1+9R}/\sqrt{1+R}$.
\begin{lemma}\label{lem:yz_triang}
Let $L(y)$ and let $M(z)$ be related by $L(y)=M(g(y)^2/y)$. Let
$\widetilde{L}(\ensuremath{\widetilde{Q}}\xspace)$ and $\widetilde{M}(\ensuremath{\widetilde{R}}\xspace)$ be the expressions of
$L(y)$ and $M(z)$ in terms of $\ensuremath{\widetilde{Q}}\xspace\equiv \ensuremath{\widetilde{Q}}\xspace(y)$ and $\ensuremath{\widetilde{R}}\xspace\equiv
\ensuremath{\widetilde{R}}\xspace(z)$, i.e., $L(y)=\widetilde{L}(\ensuremath{\widetilde{Q}}\xspace(y))$ and
$M(z)=\widetilde{M}(\ensuremath{\widetilde{R}}\xspace(z))$. Then the expressions are the same,
i.e.,
\[
\widetilde{M}=\widetilde{L}.
\]
\end{lemma}
\begin{proof}
The change of variable relation is $z=g(y)^2/y$. We have
\[
z=g(y)^2/y=Q\frac{(1-2Q)^2}{(1-Q)^3}.
\]
Hence, if we write $R\equiv R(z)=Q(y)/(1-Q(y))$, we have
$z=R(1-R)^2$, so that $R=z/(1-R)^2$. In addition we have
$Q(y)=R(z)/(1+R(z))$, so that
$\ensuremath{\widetilde{Q}}\xspace(y)=(1+8Q(y))^{1/2}=\sqrt{1+9R(z)}/\sqrt{1+R(z)}=\ensuremath{\widetilde{R}}\xspace(z)$. Hence
$\widetilde{M}=\widetilde{L}$.
\end{proof}
Define (with $z$ and $y$ related by $z=g(y)^2/y$):
\[
Z_{\infty}(z):=\frac{y}{g}Y_{\infty}(y),\ Z_i(z):=\frac{y}{g}Y_i(y),\
Z(z):=Y(y),
\]
and define
\[
C_{\infty}(z)^2:=\frac{y}{g}B_{\infty}(y)^2,\
C_i(z)^2:=\frac{y}{g}B_i(y)^2.
\]
The expression of $Y_i(y)$ in terms of $Y_{\infty}(y)$ and $Y(y)$
ensures that
\[
Z_i=Z_{\infty}\frac{(1-Z^i)(1-Z^{i+2})}{(1-Z^{i+1})^2}.
\]
The expression of $B_i(y)$ in terms of $B_{\infty}(y)$ and $Y(y)$ and
Lemma~\ref{lem:yz_triang} (to replace $\ensuremath{\widetilde{Q}}\xspace$ by $\ensuremath{\widetilde{R}}\xspace$ in the
expression) ensure that
\[
C_i=C_{\infty}\cdot\Big(1-\frac{\ensuremath{\widetilde{R}}\xspace+1}{4}Z^i\frac{(1-Z)(1-Z^2)}{(1-Z^{i+1})(1-Z^{i+2})}\Big).
\]
In addition, since $\ensuremath{H^{(k)}_i}\xspace(z)=\frac{y}{g}\ensuremath{G^{(k)}_i}\xspace(y)$, we have
\[
\ensuremath{H^{(k)}_i}\xspace(z)=Z_{i+1}-Z_{i-1}+C_i\ \!\!^2-C_{i-1}\ \!\!^2.
\]
Using Lemma~\ref{lem:yz_triang} we obtain (after simplifications):
\begin{equation}
Z_{\infty}=\frac{1}{\ensuremath{\widetilde{R}}\xspace(1-R)},\ \ C_{\infty}\ \!\!^2=\frac{16\cdot R}{(\ensuremath{\widetilde{R}}\xspace+1)^2\ensuremath{\widetilde{R}}\xspace(1-R^2)}.
\end{equation}
Lemma~\ref{lem:yz_triang} and the relation $Y+1/Y+2=1/Q$ also ensure
that $Z\equiv Z(z)$ is the algebraic generating function specified by
\begin{equation}\label{eq:Ztriang}
Z+\frac{1}{Z}+1=\frac{1}{R}.
\end{equation}
To summarize we obtain:
\begin{proposition}
For each $i\geqslant 1$ and $k\geqslant 4$, the generating function $\ensuremath{H^{(k)}_i}\xspace(z)$
has the following expression (which does not depend on $k$):
\[
\ensuremath{H^{(k)}_i}\xspace(z)=Z_{i+1}-Z_{i-1}+C_i\ \!\!^2-C_{i-1}\ \!\!^2,
\]
where
\[
Z_i=Z_{\infty}\frac{(1-Z^i)(1-Z^{i+2})}{(1-Z^{i+1})^2},\ \
C_i=C_{\infty}\cdot\Big(1-\frac{\ensuremath{\widetilde{R}}\xspace+1}{4}Z^i\frac{(1-Z)(1-Z^2)}{(1-Z^{i+1})(1-Z^{i+2})}\Big).
\]
\end{proposition}
\bibliographystyle{plain}
|
\section{1. Introduction}
Tau leptons are heavy enough that their decay products can contain an on-shell spin-2 tensor meson\footnote{Hereafter, $P$, $V$, $A$ and $T$ will denote the lowest lying pseudoscalar, vector, axial and tensor mesons, respectively.} ($J^{PC}=2^{++}$, see \cite{pdg}) in the final state. Therefore, the $\tau \to TP\nu_{\tau}$ decays can provide a unique environment to study the weak-tensor-pseudoscalar vertex in the moderate energy regime. Measurements of these hadronic matrix elements will be complementary to the ones involved in the crossed related $P \to T$ weak transitions which are accessible only in the decays of heavy $D$ and $B$ mesons. The hadronic matrix elements $\langle T|J_{\mu}|B\rangle $, which are important in the calculation of semileptonic $B \to Tl\nu$ and non-leptonic $B\to PT, VT$ or $AT$ decays, have been calculated in the framework of several effective models of QCD \cite{isgw,charles,ebert,cheng1,yang,wang,cheng,others,Cheng-Chiang2010}. Manifestly, semileptonic decays provide a cleaner environment to study the weak $PT$ vertex owing to an exact factorization of the decay amplitude, while the non-leptonic amplitudes receive contributions from different terms of the effective weak hamiltonian and a factorization approximation is usually assumed in some calculations (for an extensive literature on the subject see \cite{isgw,charles,ebert,cheng1,yang,wang,cheng,others,Cheng-Chiang2010,verma,lopez,kim}). From the experimental side, a few measurements or upper limits about some of these $B$ meson decay channels have been reported so far by $B$-factory experiments (results are listed in \cite{pdg}) and a proper account of the measured rates is still the subject of current investigations. Conversely, tensor mesons produced in $\tau$ lepton decays have been scarcely investigated at the theoretical level and, from the experimental point of view, only the upper limit $B(\tau \rightarrow K_{2}^*(1430)\nu _{\tau })<
3\times 10^{-3}$ has been reported in \cite{pdg}. As it was discussed in \cite{3}, if it was observed, this decay mode would require the existence of exotic tensor charged weak currents.
As is well known, measurements of $\tau $ decays involving
two or more pseudoscalar mesons have shown the presence of several intermediate
resonant states which populate the different hadronic invariant-mass spectra \cite{Davier:2005xq, Lee:2010tc, Aubert:2007mh}. Indeed, these hadronic spectra have been useful to determine the properties of $\rho,\ \rho',\ a_1(1260)$ and $K^*$ resonances in a clean environment (see discussion in \cite{Davier:2005xq}). Recently, both BaBar and Belle collaborations have reported refined measurements of $\tau$ decays into three pseudoscalar mesons which include either pions and/or kaons \cite{Lee:2010tc, Aubert:2007mh}. Since tensor mesons undergo sizable decay rates to two pseudoscalar mesons in a $d$-wave orbital configuration \cite{pdg}, one may expect that $T$ mesons give a contribution to three-pseudoscalar $\tau$ lepton decays via the $\tau \to P_1T^*(\to P_2P_3)\nu_{\tau}$ decay chain; eventually, we would be able to extract the $\tau \to TP\nu_{\tau}$ rates from the relevant hadronic spectra as it was done recently to extract the branching fractions for the $\tau^- \to \phi \pi^-\nu, \phi K^-\nu, KK^*\nu$ \cite{Lee:2010tc, Aubert:2007mh} decay modes from data on the three-pseudoscalar channels of tau decays.
In this paper we study the $TP$ channels that are kinematically allowed in $\tau$ lepton decays. Most of the popular effective models of QCD at low energies do not make predictions for the weak $PT$ vertex in the energy region relevant for $\tau$ decays. Here we use a meson dominance model where the weak and strong coupling constants are determined from other independent decay processes (see for example \cite{Gabriel2008} for an application to $\tau \to (\omega, \phi)\pi \nu$ decays). We find that the branching fractions for the $\tau \to T\pi\nu$ channels under study are of the order of $10^{-6}\sim 10^{-4}$ and therefore, intermediate tensor resonances give a small contribution to the rates of three-pseudoscalar final states. Eventually, the large data sample of $\tau^+\tau^-$ pairs accumulated by $B$-factory experiments would allow to extract the rates of tensor mesons produced in $\tau$ lepton decays.
\bigskip
\section{2. A meson dominance model for $\protect\tau \rightarrow
T \pi \nu $ decays.}
\bigskip
Let us consider the $\tau^-
(p_{1})\rightarrow T(p_{T})\pi (p)\nu_{\tau} (p_{2})$ decay, where $T$ denotes an on-shell tensor meson; analogous decays involving a kaon or $\eta$ meson are (almost) forbidden by kinematics. The decay amplitude for this process is given by:
\begin{equation}
{\cal M}(\tau \rightarrow T\pi \nu
)=\frac{G_{F}}{\sqrt{2}}V_{uD}\bar{u}%
(p_{2},s_{2})\gamma _{\mu }(1-\gamma _{5})u(p_{1},s_{1})\langle T(p_{T})\pi (p)|J^{\mu }(0)|0\rangle , \label{1}
\end{equation}
where $V_{uD}$ ($D=d$ or $s$) is the $uD$ entry of the Cabibbo-Kobayashi-Maskawa matrix and $J_{\mu}(0)$ is the corresponding $V-A$ weak current.
The hadronic matrix element $\langle T\pi |J^{\mu }|0\rangle$
can be parametrized as follows \cite{isgw}:
\begin{eqnarray}
\langle T(p_{T},\varepsilon )\pi (p)|J^{\mu }|0\rangle
&=&ih\epsilon ^{\mu \nu \phi \rho }\varepsilon _{\nu \alpha }^{\ast
}p^{\alpha }(p-p_{T})_{\phi }(p+p_{T})_{\rho }-k\varepsilon ^{\ast \mu \nu
}p_{\nu } \nonumber \\
&&\ -\varepsilon ^{\ast \alpha \beta }p_{\alpha }p_{\beta }[b_{+}(p-p_{T})^{\mu
}+b_{-}(p+p_{T})^{\mu }], \label{2}
\end{eqnarray}
where $h(t),$ $k(t)$, $b_{\pm }(t)$ are Lorentz-invariant form factors and $t=(p_T+p)^{2}$ is the square of the momentum transfer. The symmetric tensor $\varepsilon^{\ast \mu\nu}$ describes the spin-2 polarization states of the outgoing tensor meson.
The unpolarized squared amplitude becomes:
\begin{equation}
\sum_{pols}|{\cal M}|^{2}=2G_{F}^{2}|V_{uD}|^{2}\left[
c_{1}(u,t)|h|^{2}+c_{2}(u,t)|k|^{2}+c_{3}(u,t)|b_{-}|^{2}+c_{4}(u,t) {\rm
Re}(k^{\ast }b_{-})\right]\ , \label{3}
\end{equation}
where we have defined $u=(p_1-p)^2$, and $c_i=c_i(u,t)$ are kinematical
factors given by:
\begin{eqnarray}
c_1&=&\frac{\lambda }{8m_{T}^{2}}\left\{2tu^{2}+2\left[ m_{\tau
}^{2}(m_{\pi
}^{2}-m_{T}^{2})-t(m_{T}^{2}+m_{\tau }^{2}+m_{\pi }^{2}-t)\right] u
\right. \nonumber \\ && \left. +\frac{1}{2}(m_{\tau }^{2}-t) \left[
-2\lambda +t(2m_{\pi
}^{2}+2m_{T}^{2}-t)-m_{\tau }^{2}(m_{\tau }^{2}-t)-6m_{T}^{2}m_{\pi
}^{2}\right] \right. \nonumber \\
&& \left. +\frac{1}{2}m_{\tau }^{2}(2m_{T}^{2}+m_{\tau
}^{2}-t)^{2}\right\},
\\ \label{4}
c_{2}&=& \frac{1}{12m_{T}^{4}}\left\{(\lambda +tm_{T}^{2})u^{2}+\left[
(t-m_{\pi}^{2}-m_{T}^{2})(\lambda +m_{T}^{2}t)-m_{T}^{2}m_{\tau
}^{2}(t+m_{T}^{2}-m_{\pi }^{2})\right]u \right. \nonumber \\
&& \left. +\frac{1}{2}m_{T}^{2}(m_{\tau }^{2}+2m_{\pi }^{2}-3t)\lambda
+m_{T}^{4}m_{\tau }^{2}(m_{\tau }^{2}+m_{T}^{2}-t)-m_{\pi
}^{2}m_{T}^{4}(m_{\tau }^{2}-t)\right\}, \\ \label{5}
c_{3}&=& \frac{m_{\tau }^{2}(m_{\tau }^{2}-t)}{48m_{T}^{4}}\lambda ^{2},
\\ \label{6}
c_{4}&=& \frac{m_{\tau }^{2}\lambda
}{12m_{T}^{4}}\left\{(t+m_{T}^{2}-m_{\pi
}^{2})u-m_{T}^{2}(2m_{\tau }^{2}+m_{T}^{2}-m_{\pi }^{2}-t)\right\}.
\label{7}
\end{eqnarray}
where $\lambda = t^2 + m_T^4+m_{\pi}^4-2tm_{T}^2-2tm_{\pi}^2-2m_T^2m_{\pi}^2$.
As we have pointed out previously, we resort to a meson dominance model to compute the form factors in our $\tau$ decays (see Figure \ref{fig1}). For definiteness, we will illustrate the method in the case of the $\tau \rightarrow K_{2}^{\ast }(1430)\pi \nu $ decay, because in this case all form factors receive contributions from intermediate $t$-channel virtual states.
In this model we will assume that the above decay receives contributions from three intermediate states: the pseudoscalar $K$ and axial $K_{1}(1400)$ mesons which saturate the axial current, and the vector meson $K^{\ast }(892)$ which contributes to the vector current. Other meson resonances can
contribute as well to both currents; we would expect their corrections to be small since either their strong couplings to the $K_2^*\pi$ system or their couplings to the weak current are suppressed. Such additional contributions may be enhanced if their resonance shapes were peaked in the kinematical domain of $\tau$ decays ($(m_T+m_{\pi})^2 \leq t \leq m_{\tau}^2$), which is not the case.
Within our approximations, the decay amplitude is given by (see Figure 1)
\begin{equation}
{\cal M} (\tau \rightarrow K_{2}^{\ast }\pi \nu )=\sum_{R=K, \ K^*(892), \ K_1(1400) }\!\!\!\!\!\!\!\!\!\!{\cal M}(\tau
\rightarrow R\nu
\rightarrow K_{2}^{\ast }\pi \nu )\ .
\label{8}
\end{equation}
\begin{figure}
\includegraphics[width=18cm]{Figure1.eps}
\vspace{-18.0cm}
\caption{Intermediate meson dominance graph of $\tau \to T\pi\nu$ decays}\label{fig1}
\end{figure}
Using the Feynman rules to compute the above amplitudes and comparing the results with Eq. (\ref{2}), we derive the following expressions
for the form factors:
\begin{eqnarray}
k(t)&=&- \frac{if_{K_{1}}g_{K_{2}^{\ast }K_{1}\pi
}}{m_{K_1}}\cdot {\rm BW}_{K_1}(t), \\ \label{9}
h(t)&=&\frac{f_{K^{\ast }}g_{K_{2}^{\ast }K^{\ast }\pi
}}{2m_{K^*}}\cdot {\rm BW}_{K^*}(t),
\\ \label{10}
b_{-}(t)&=&\frac{if_{K_{1}}g_{K_{2}^{\ast }K_{1}\pi }}{m_{K_{1}}^3}\cdot {\rm BW}_{K_1}(t)+
\frac{f_{K}g_{K_{2}^{\ast }K\pi }}{m_K^2}\cdot {\rm BW}_K(t), \\
\label{11} b_{+}(t)&=& 0, \label{12}
\end{eqnarray}
where $f_X$ and $g_{K_2^* X\pi}$ denote the weak and strong coupling constants of
intermediate $X$ states. The Breit-Wigner forms introduced above are defined as ${\rm BW}_X(t)=m_X^2/(m_X^2-t-im_X\Gamma_X \theta(t-t_{\rm threshold}))$, with $m_X$ and $\Gamma_X$ being the mass and decay width (which we choose to be a constant) of the resonance.
In a similar way, we can assume the same meson dominance model to describe the
strangeness-conserving $\tau \rightarrow a_{2}(1320)\pi \nu$ decay.
Owing to G-parity conservation \cite{scc}, the amplitude for the $\tau \rightarrow a_{2}\pi \nu$ process will receive
contributions only in the vector current via the following decay chain $\tau \to
(\rho(770), \rho'(1450))\nu \to a_2(1320)\pi\nu$, where $\rho'$ is the first radial excitation of the $\rho$ meson. In this case, the only non-vanishing form factor becomes:
\begin{equation}
h(t)=\ \frac{f_{\rho }g_{a_{2}\rho \pi }}{2m_{\rho}}\cdot \frac{{\rm BW}_{\rho}(t)+\beta {\rm BW}_{\rho'}(t) }{1+\beta}\ , \label{13}
\end{equation}
where $\beta$ denotes the ratio of $\rho'$ to $\rho$ coupling constants, and it is similar to one defined in the two-pseudoscalar decay modes (see for example \cite{beta}). In the studies of these two-pseudoscalar decays of $\tau$ leptons carried out by ALEPH \cite{taupipi} and Belle \cite{taukpi} Collaborations, $\beta$ turns out to be small and almost real: $\beta_{\pi\pi} \approx -0.15 $ \cite{taupipi} and $\beta_{K\pi} \approx 0.08$ \cite{taukpi}, for the $\pi^-\pi^0$ and $K^0\pi^-$ decay modes, respectively. In our present calculation, we will assume $\beta_{a_2\pi}=(\pm 0.2 \pm 0.1)$ as a rather conservative value.
Finally, we also consider the $\tau \rightarrow f_{2}(1270)\pi \nu$ decay. In this case $G$-parity conservation forbids the contribution of the vector current to the decay amplitude. We assume that the dominant contributions come from the pseudoscalar and axial resonances by means of the chain $\tau \to (\pi,\ a_1) \nu \to f_2 \pi \nu$. We also assume that the mixing angle between the $f_2(1270)$ and $f_2'(1525)$ tensor mesons is such that the $f_2$ is dominantly a $(u\overline{u} + d\overline{d})/\sqrt{2}$ state \cite{Cheng-Chiang2010}. The only non-vanishing form factors in this case become:
\begin{eqnarray}
k(t)&=&- \frac{if_{a_{1}}g_{f_{2}a_{1}\pi
}}{m_{a_1}}\cdot {\rm BW}_{a_1}(t), \\ \label{14}
b_{-}(t)&=&\frac{if_{a_{1}}g_{f_{2}a_{1}\pi }}{m_{a_{1}}^3}\cdot {\rm BW}_{a_1}(t)+
\frac{f_{\pi}g_{f_{2}\pi\pi }}{m_{\pi}^2}\cdot {\rm BW}_{\pi}(t), \label{15}
\end{eqnarray}
\bigskip
\section{3. Determination of the strong and weak couplings.}
\bigskip
In this section we focus on the determination of the strong and weak coupling
constants that appear in Eqs.(9-15). We first consider in more detail
the decay widths of the $T\rightarrow P\pi,\ V\pi, \ A\pi$,
and $A\rightarrow V\pi$ decays reported in Ref. \cite{pdg} to determine the strong couplings.
The decay constants for the above decays are defined from the following decay amplitudes \cite{7}, which assume that only one single $L$-wave configuration contributes to the final states:
\begin{eqnarray}
{\cal M}[T(p_{T},\varepsilon )\longrightarrow P(p_{P})\pi (p)]&=&
g_{TP\pi}\varepsilon ^{\mu \alpha }p_{\mu }p_{\alpha } \\ \label{14}
{\cal M} [T(p_{T},\varepsilon )\longrightarrow V(\epsilon ,p_{V})\pi
(p)]&=& g_{TV\pi }\epsilon _{\mu \nu \rho \sigma }\epsilon ^{\mu
}p_{V}^{\nu }\varepsilon^{\rho \alpha }p_{\alpha }p^{\sigma } \\
\label{15}
{\cal M}[T(p_{T},\varepsilon )\longrightarrow A(\epsilon ,p_{A})\pi
(p)]&=& g_{TA\pi }\varepsilon _{\alpha \beta }\epsilon ^{\alpha }p^{\beta
} \\ \label{16}
{\cal M}[A(\epsilon _{A},p_{A})\longrightarrow V(\epsilon
_{V},p_{V})P(p)]&=& g_{AVP}(\epsilon _{A})_{\mu }(\epsilon _{V})_{\nu
}\left\{ p_{A}.p_{V}g^{\mu \nu }-p_{A}^{\nu }p_{V}^{\mu }\right\} .
\label{17}
\end{eqnarray}
The corresponding decay rates in the rest frame of the decaying particle are:
\begin{eqnarray}
\Gamma (T\longrightarrow P\pi )&=&\frac{g_{TP\pi }^{2}}{60\pi m_{T}^{2}}%
\left\vert \vec{P}_{c}\right\vert ^{5}\ , \\ \label{18}
\Gamma (T\longrightarrow V\pi )&=&\frac{g_{TV\pi }^{2}}{40\pi }\left\vert
\vec{P}_{c}\right\vert ^{5}\ , \\ \label{19}
\Gamma (T\longrightarrow A\pi )&=& \frac{g_{TA\pi }^{2}}{120\pi m_{T}^{2}m_{A}^{2}}\left( 2\left\vert
\vec{P}_{c}\right\vert ^{5}+5m_{A}^{2}\left\vert \vec{P}_{c}
\right\vert^{3}\right) \ , \\ \label{20}
\Gamma (A \rightarrow VP)&=& \frac{g_{AVP}^{2}}{8\pi m_{A}^{2}m_{V}^{2}}
\left[m_{A}^{2}|\vec{P}_{c}|^{5}-\frac{1}{4}(m_{A}^{2}+m_{V}^{2}
-m_{P}^{2})^{2}| \vec{P}_{c}|^{3}\right. \nonumber \\
&& \left. +\frac{3}{4}m_{V}^{2}(m_{A}^{2}+m_{V}^{2}-m_{P}^{2})^{2}|
\vec{P}_{c}|\right]\ ,
\end{eqnarray}
where $\vec{P}_c$ denotes the three-momentum of anyone of the particles in the final state.
In order to extract the decay constant $g_{K_{2}^{\ast }K_{1}\pi }$ we
have assumed that the experimentally measured rate of $K_{2}^{\ast
}(p_{T})\rightarrow K^{\ast }(p_{V})\pi (p_{2})\pi (p_{1})$ is
saturated by the contribution of the $K_1(1400)$ intermediate state
through the chain process $K_{2}^{\ast }\rightarrow K_{1}(1400)\pi
\rightarrow K^{\ast }\pi \pi $. The dominance of this mechanism is also assumed in other works (see for example \cite{axial}). Of course, we are aware that other intermediate resonances might also contribute (for example the $\rho$, $\sigma$ and $f_0$ resonances in the $\pi\pi$ channel), but either their couplings to
$K_2^*K^*$ are small or forbidden (an alternative view of the problem is discussed in Ref. \cite{singer}, which considers that the dominant contribution arises from the $g_{TVV}$ coupling).
We assume isospin symmetry to relate the strong coupling constants for different charge states in a given channel, and SU(3) flavor symmetry to relate the couplings of vertices that can not be measured directly to the ones that are extracted from measured rates. Using these approximations and the measured rates \cite{pdg} of relevant decays, we obtain the following central values:
$g_{K_{2}^{\ast -}K^{-}\pi ^{0}}=8.35$ GeV$^{-1}$, $g_{K_{2}^{\ast
-}K^{\ast -}\pi ^{0}}=9.02$ GeV$^{-2},$ $g_{K_{1}^{-}K^{\ast -}\pi
^{0}}=1.95$ GeV$^{-1}$ , $g_{K_{2}^{\ast -}K_{1}^{-}\pi ^{0}}=30.6$,
$\ g_{\overline{K_{2}^{\ast 0}}K_{1}^{-}\pi^+}=-43.3$, $g_{f_2a_1^+\pi^-}=32.3$,
$g_{\overline{K_{2}^{\ast 0}}K^{-}\pi^+}=12.94$ $\ $GeV$^{-1},\ g_{\overline{K_{2}^{\ast 0}}K^{\ast -}\pi ^{+}}=12.26$ GeV$^{-2},\
g_{a_{2}^{-}\rho ^{-}\pi ^{0}}=19.40$ GeV$^{-2}$, $g_{a_{2}^{0}\rho
^{-}\pi ^{+}}=19.47$ GeV$^{-2}$, and $g_{f_2\pi^+\pi^-}=20.3$ GeV$^{-1}$. In order to get some of the couplings involving the strange axial mesons \cite{Gabriel2008} we have assumed a value of $\theta_A=50.8^{\circ}$ \cite{Cheng-Chiang2010} for the mixing angle of the $K_1(1270)$ and $K_1(1400)$ strange mesons. We also note that the value of the $f_2\pi\pi$ coupling given above agrees with the prediction obtained in the Appendix of the first paper in Ref. \cite{singer}. We note that the uncertainties associated to these couplings are estimated directly from their measured masses and rates or, when data was not available, they were attributed a conservative $20\%$ uncertainty if SU(3) symmetry was assumed in their derivation.
Finally, the values of other relevant inputs to determine the branching ratios
of tau decays have been taken from Ref. \cite{pdg}. In addition we have
set the weak coupling of hadron $H^-$ from $\tau^- \to
H^-\nu$ decays: $f_{\pi}=(130.7\pm 0.4)$ MeV, $f_{K^{-}}=(159.8\pm 1.5)$ MeV, $f_{K^{\ast -}}=(210\pm 5)$ MeV, and
$f_{\rho }=(218\pm 2)$ MeV. On the other hand, we use $f_{a_1}= (238\pm 10)$ MeV in our calculations and we have taken $f_{K_1(1400)}= (-139^{+41}_{-46})$ MeV from Ref. \cite{Cheng-Chiang2010}.
\section{4. Branching ratios of tau decays into tensor mesons.}
The branching ratios predicted in this work for the $\tau^- \to T\pi\nu$ decays are shown in Table 1. The main uncertainty in the rates of Cabibbo-suppressed channels comes from the large error bar in the $f_{K_1}$ coupling constant, while the one in the $a_2\pi$ channel is dominated by the large uncertainty ($\pm 50\%$) that we have attributed to the value of $\beta$. Since our model uses Breit-Wigner forms with a constant decay width and given that the contributions of higher mass virtual states have been neglected in our calculation of the form factors, further uncertainties are expected to contribute to the results shown in Table 1.
In addition, in our model we have not considered the contribution of the continuum which can be associated to a contact (non-BW) term in the weak $T\pi$ vertex. We have not estimated these uncertainties in Table 1; eventually, the continuum contributions may be large, but they are rather difficult to evaluate in the meson dominance model like ours in the absence of constraints about such contact terms.
\begin{table}[t]
\begin{tabular}{|c|c|c|}
\hline
\ \ \ $T\pi$ mode & \ \ \ \ \ \ Branching ratio & \ \ \ \ \ \ Comment\ \ \ \\
\hline \hline
$K_2^{*-}(1430)\pi^0$ & $(3.7\pm 2.1)\times 10^{-6}$ & \\
\hline
$\overline{K}_2^{*0}(1430)\pi^-$ &$(4.7\pm 2.7) \times 10^{-6}$ & \\
\hline
$a_2^-(1320)\pi^0$ & $(9.4\pm 4.8)\times 10^{-6}$ & $\beta=+0.2\pm 0.1$,\\
& $(10.1\pm 7.5)\times 10^{-6}$ & $\beta=-0.2\pm 0.1$ \\
\hline
$a_2^0(1320)\pi^-$ & $(9.0\pm 4.6)\times 10^{-6}$ & $\beta=+0.2\pm 0.1$,\\
& $(9.5\pm 7.1)\times 10^{-6}$ & $\beta=-0.2\pm 0.1$ \\
\hline
$f_2(1270)\pi^-$ & $(5.9\pm 1.8)\times 10^{-4}$ & $f_2$ pure $\bar{u}u,\ \bar{d}d$ \\
\hline
\end{tabular}
\caption{Branching ratios for the $(T\pi)$ decays of the $\tau$ lepton.}
\end{table}
The branching fractions turn out to be of order $10^{-4}\sim 10^{-6}$, with the largest rate corresponding to the $f_2(1270)\pi^-$ decay mode. Therefore, we can expect that the contribution of tensor meson intermediate states to the three-pseudoscalar decays of tau leptons is small. Concerning the results shown in Table 1, we observe that the Cabibbo-favored decay involving the $f_2$ meson is larger that the one involving the $a_2$ meson because, owing to $G$-parity, the former receives contributions from the dominant axial current while the second is mediated by the vector current only. Similarly, the Cabibbo-suppressed channels are of similar size as the Cabibbo-allowed $a_2\pi$ decays because the former receive contributions of the vector and axial currents. In addition to the above dynamical considerations, we should point out that these $T\pi$ channels are suppressed mainly due to the reduced phase space available in $\tau$ lepton decays.
\bigskip
\bigskip
\section{5. Conclusions}
We have studied and computed the branching ratios of the $\tau \rightarrow (K_{2}^{\ast }, a_{2}, f_2)\pi \nu $ decays; we have not considered final states involving kaons or $\eta$ mesons because either they are suppressed or forbidden by kinematics. We have used a meson dominance model where the form factors are dominated by the lowest lying resonances that couple to the $T\pi$ system. To our knowledge, this is the first study reporting results on these peculiar $\tau$ lepton decays. Beyond probing the tensor-pseudoscalar weak vertex, the processes under consideration can contribute as intermediate states in $\tau$ lepton decays involving three-pseudoscalar mesons.
Owing to $G$-parity of strong interactions, the rates of the Cabibbo allowed and suppressed decay channels exhibit an interesting pattern. The Cabibbo-supressed $K_2^*\pi$ channels turn our to be of the same order as the Cabibbo-favored $a_2\pi$ decays, mainly because the latter receives contributions only from the vector current. The calculated branching fractions spread from $10^{-4}$ to $10^{-6}$, with the largest branching fraction ($\approx 6\times 10^{-4}$) corresponding to the $f_2(1270)\pi^-$ final state. Eventually, these decays will be measured from the invariant mass distributions of the decay products of the intermediate $T$ tensor mesons in three-body decays of $\tau$ leptons, given the large data sample of $\tau$ lepton pairs recorded by the Babar and Belle experiments \cite{9}.
\bigskip
\subsection{Acknowledgements}
GLC acknowledges financial support
from Conacyt and SNI (M\'{e}xico). JHM is grateful to \textit{%
Comit\'{e} Central de Investigaciones} (CCI) of the University of Tolima and CNPq (Brazil) for
financial support and also thanks to the Physics Department at Cinvestav for
the hospitality while this work was ended.
\bigskip
|
\section{\label{sec:Intro}Introduction}
Within the last decades, it has become more and more apparent that neutrinos can be used as one of the major windows to look for physics beyond the Standard Model (SM) of elementary particles, which is often dubbed ``new physics''. Not only has it been established by neutrino oscillation experiments on solar~\cite{Abdurashitov:2002nt,Altmann:2005ix,Aharmim:2007nv,Arpesella:2007xf}, atmospheric~\cite{Ashie:2005ik}, and artificially produced neutrinos~\cite{KamLAND:2008ee,Ahn:2006zza,Adamson:2007gu} that neutrinos are massive, but also the corresponding leptonic mixing differs significantly from the quark sector~\cite{Nakamura:2010zzi}. Apart from their values, one of the major questions about neutrino masses is their origin. A particularly attractive possibility to explain their smallness is the so-called ``seesaw mechanism''~\cite{Minkowski:1977sc,Yanagida:1979as,Glashow:1979nm,GellMann:1980vs,Mohapatra:1979ia}. This mechanism, however, intrinsically involves the breaking of lepton number by two units. If this breaking was indeed realized in Nature, a process that would have to appear is neutrino-less double beta decay ($0\nu\beta\beta$), where an even-even nucleus undergoes a transition $(A,Z)\to (A,Z+2) + 2 e^-$, thereby violating lepton number. This process has been intensively searched for, but up to now there has not yet been an unambiguous detection~\cite{Aalseth:2004hb} (cf., however, the claim by part of the Heidelberg-Moscow collaboration in Ref.~\cite{KlapdorKleingrothaus:2004wj}).
Although the ``standard'' mechanism for \mbox{$0\nu\beta\beta$}\ is the exchange of light Majorana neutrinos~\cite{Doi:1985dx,Bilenky:1987ty,Schmitz:1997ag}, other mechanisms could very well appear in certain models beyond the SM, such as supersymmetry (see, e.g., Refs.~\cite{Mohapatra:1986su,Allanach:2009iv}) or left-right symmetric models (see, e.g., Refs.~\cite{Mohapatra:1986pj,Tello:2010am}). To disentangle the various possibilities, it will probably be necessary to detect the process for several different isotopes~\cite{Deppisch:2006hb,Gehman:2007qg}.
In this paper, we will investigate point-like contributions of new physics to \mbox{$0\nu\beta\beta$}, previously studied in Refs.~\cite{Pas:2000vn,Prezeau:2003xn}. Over the last years, there has been considerable improvement in the knowledge of the neutrino oscillation parameters~\cite{Schwetz:2011qt}, and new experiments on neutrino masses and mixings are currently starting or will start in the near future. One of the most prominent near future experiments on neutrino-less double beta decay is certainly the GERDA experiment~\cite{Abt:2004yk}, which we will take as an example. This experiment will be complemented by others aiming to detect the kinematical electron-neutrino effective mass $m_\beta$ (e.g.\ KATRIN~\cite{Osipowicz:2001sq}), the cosmological sum $\Sigma$ of neutrino masses (e.g.\ Planck~\cite{Planck:2011ah}), or the leptonic mixing angle $\theta_{13}$ (e.g.\ Double Chooz~\cite{Ardellier:2004ui}). We therefore feel that it is necessary to investigate the impact of these experiments on new physics contributions to \mbox{$0\nu\beta\beta$}. Naturally, once this process is observed, there will soon be analyses of the standard picture. Later on, however, the question will arise what else we could learn from a positive or negative signal of \mbox{$0\nu\beta\beta$}. Here, we are going to give an answer to this question in what regards the short-range contributions to the process.
The paper is organized as follows: In Secs.~\ref{sec:EFT} and~\ref{sec:chi2}, we will shortly review the nature of the short-range contributions to \mbox{$0\nu\beta\beta$}, as well as the analysis of a signal under the assumption of the standard mechanism of light neutrino exchange. Our actual investigations will be presented in Secs.~\ref{sec:scenarios} and~\ref{sec:Rates}, where we first define certain benchmark scenarios and use them later on to illustrate how strong the bounds on the effective operators of the new physics contributions to \mbox{$0\nu\beta\beta$}\ could be for certain realistic situations after the future experiments have yielded results. Depending on if the signal of \mbox{$0\nu\beta\beta$}\ is positive or negative, and also depending on if it is consistent or inconsistent with results from complementary experiments, the bounds can be stronger or weaker. In general, it is hard to derive strict statistical bounds, due to the intrinsic difficulties arising from the nuclear physics involved. Finally, we present our conclusions in Sec.~\ref{sec:Conc}.
\section{\label{sec:EFT}Point-like contributions to $\boldsymbol{0\nu\beta\beta}$}
The key point of our analysis is that one can parametrize many different contributions of new physics to $0\nu\beta\beta$ in a nearly model-independent way by using effective field theory~\cite{Georgi:1994qn,Pich:1998xt,Manohar:1996cq}. This is a very powerful approach as long as the detailed spectrum of new heavy particles does not play a significant role.\footnote{See, e.g., Ref.~\cite{Blennow:2010th} for a case where the detailed heavy spectrum {\it does} play a role.} Depending on if one only modifies the vertices that couple to the light neutrinos or if the whole decay mechanism is altered, one obtains long-range~\cite{Pas:1997fx} and short-range~\cite{Pas:2000vn} contributions, respectively. Here, we restrict ourselves to the latter ones, since modifications of the vertices only are even discussed in some textbooks, see e.g.\ Ref.~\cite{Schmitz:1997ag}, and furthermore such modified weak interaction charged current vertices can be probed much more easily in, e.g., neutron beta decay experiments~\cite{Severijns:2006dr} or in muon conversion experiments~\cite{LeeRoberts:2007gf}.
When looking at the short range contributions, one has to study the following general Lagrangian~\cite{Pas:2000vn}:
\begin{equation}
\mathcal{L}=\frac{G_F^2}{2}m_p^{-1} \left( \epsilon_1 JJj + \epsilon_2 J^{\mu\nu} J_{\mu\nu} j + \epsilon_3 J^\mu J_\mu j + \epsilon_4 J^\mu J_{\mu\nu} j^\nu + \epsilon_5 J^\mu J j_\mu \right) + \text{h.c.}
\label{eq:L_0nbb}
\end{equation}
Here, $G_F$ is the Fermi constant, which is a reflection of the double weak interaction vertex involved in the standard light neutrino exchange diagram for \mbox{$0\nu\beta\beta$}, and $m_p$ is the proton mass, which is used to give the Lagrangian the correct mass dimension. In principle, any high energy scale could have served this purpose, but a scale of the order of the proton mass is already considerably higher than the typical internal momentum in the nucleus, which is of the order of 100~MeV~\cite{Bilenky:1987ty}. The strengths of the different operators are parametrized by the (generally complex) dimensionless coefficients $\epsilon_i$. The hadronic currents in Eq.~\eqref{eq:L_0nbb} are given by
\begin{equation}
J_{L,R} = \overline{u}\left( 1 \mp \gamma_5 \right) d, \quad J^\mu_{L,R} = \overline{u} \gamma^\mu \left( 1 \mp \gamma_5 \right) d, \quad J^{\mu\nu}_{L,R} = \overline{u} \frac{i}{2} \left[\gamma^\mu, \gamma^\nu\right] \left( 1 \mp \gamma_5 \right) d,
\label{eq:hadron_operators}
\end{equation}
and the leptonic ones are
\begin{equation}
j_{L,R} = \overline{e}\left( 1 \mp \gamma_5 \right) e^c = 2\ \overline{e_{R,L}} (e_{R,L})^c, \quad j^\mu_{L,R} = \overline{e} \gamma^\mu \left( 1 \mp \gamma_5 \right) e^c = 2\ \overline{e_{L,R}} \gamma^\mu (e_{R,L})^c.
\label{eq:electron_operators}
\end{equation}
For all currents, different chirality structures are permitted. Note that we have also assigned the correct chirality to the electron operators in Eq.~\eqref{eq:electron_operators}, since we will need this for our argumentation later on.
Actually, we could have added some more Lorentz-invariant terms to Eq.~\eqref{eq:L_0nbb},
\begin{equation}
\mathcal{L}'=\frac{G_F^2}{2}m_p^{-1} \left( \epsilon_6 J^\mu J^\nu j_{\mu\nu} + \epsilon_7 J J^{\mu\nu} j_{\mu\nu} + \epsilon_8 J_{\mu\alpha} J^{\nu\alpha} j_\nu^\mu \right),
\label{eq:Lprime_0nbb}
\end{equation}
where the leptonic tensor current is given by $j^{\mu\nu}_{L,R} = \overline{e} \frac{i}{2} \left[\gamma^\mu, \gamma^\nu\right] \left( 1 \mp \gamma_5 \right) e^c$. These terms had been included in the Lagrangian in Ref.~\cite{Pas:2000vn}, but were neglected in the final analysis, since these authors worked in the $s$-wave approximation where such contributions vanish. In Ref.~\cite{Prezeau:2003xn}, however, it was pointed out that all operators proportional to $\overline{e}\gamma_\mu e^c$, $\overline{e}\frac{i}{2}\left[\gamma_\mu,\gamma_\nu\right] e^c$, and $\overline{e}\gamma_5 \frac{i}{2}\left[\gamma_\mu,\gamma_\nu\right] e^c$ vanish identically, since the electron fields are Grassmann numbers. Therefore, the terms in Eq.~\eqref{eq:Lprime_0nbb} are not relevant for \mbox{$0\nu\beta\beta$}.
The purpose of the effective field theory approach is to provide an easy and unified description of new physics. In particular, since there is only a finite set of operators, it allows to efficiently compare different underlying models in what regards their interaction strengths. Very recent works covering different explicit $0\nu\beta\beta$ mechanisms and in particular their interplay with the standard light neutrino exchange~\cite{Faessler:2011qw,Faessler:2011rv} appeared nearly simultaneously to this paper. All the mechanisms treated there can be translated into our effective language.
To give one explicit and illustrative example, we take the gluino exchange mechanism in $R$-parity violating supersymmetry (RPV-SUSY)~\cite{Faessler:2011qw,Faessler:2011rv,Chemtob:2004xr}. The goal is to reproduce Eq.~\eqref{eq:L_0nbb} by integrating out all heavy particles, and it is easy to see that the corresponding effective Lagrangian is given by
\begin{equation}
\mathcal{L}_{\rm RPV}=\frac{G_F^2}{2}m_p^{-1} (\epsilon^{RRR}_1)_{\rm RPV}\ J_R J_R j_R + \text{h.c.},
\label{eq:L_RPV}
\end{equation}
whose coefficient can be expressed in terms of model parameters as~\cite{Faessler:2011qw}
\begin{equation}
(\epsilon^{RRR}_1)_{\rm RPV} = \frac{\pi \alpha_s}{6} \frac{(\lambda'_{111})^2}{G_F^2 m_{\tilde d_R}^4} \frac{m_p}{m_{\tilde g}} \left[ 1+ \left( \frac{m_{\tilde d_R}}{m_{\tilde u_L}} \right)^2 \right]^2,
\label{eq:eps_RPV}
\end{equation}
where $\alpha_s=g_3^2/(4\pi)$ is the strong fine structure constant, $m_{\tilde u_L,\tilde d_R,\tilde g}$ are the masses of the up-squark, of the down-squark, and of the gluino, respectively, and the parameter $\lambda'_{111}$ encodes the strength of the RPV-SUSY operators. Obviously, if we have any kind of bound on the general coefficient $\epsilon^{RRR}_1$, this can be translated into a bound on the parameter $\lambda'_{111}$, and hence on the size of the new physics contribution.
Later on, we will derive bounds on $\lambda'_{111}$ for all scenarios under consideration, in order to illustrate how one can translate our results into bounds on concrete models of new physics. In order to arrive at numerical values for $\lambda'_{111}$, we will assume that $m_{\tilde u_L}=m_{\tilde d_R}=800$~GeV and $m_{\tilde g}=1000$~GeV. These values are good examples for the Minimal Supersymmetric Standard Model squark and gluino masses at low energies, when a minimal Supergravity scenario is considered to be present at the GUT scale~\cite{Martin:1997ns}. A generalization of this procedure to underlying models different from the RPV-SUSY example is straightforward.
In Sec.~\ref{sec:Rates}, after having introduced the standard mechanism, we will discuss the actual formula for the corresponding decay rate.
\section{\label{sec:chi2}The standard analysis of a positive signal}
The most interesting case to study is the one where the GERDA experiment yields a positive signal. This corresponds to a measured decay rate $\Gamma_{\rm obs}$ that is different from zero and has a standard deviation $\sigma_{\rm obs}$:\footnote{Note that, depending on the size of $\sigma_{\rm obs}$, it is possible that this rate is also consistent with a zero decay rate at $x\cdot \sigma$ if $\Gamma_{\rm obs}- x \cdot \sigma_{\rm obs}=0$. We will ignore this subtlety here and assume that a positive experimental signal means that lepton number violation is experimentally established. This is a reasonable assumption in the likely case that we trust the experimentalists conducting the experiment. Furthermore, the analysis would become much more subtle when including such pathologic cases, without any significant gain.}
\begin{equation}
\Gamma_{\rm obs}\pm \sigma_{\rm obs}\ @~68~\%~{\rm C.L.}
\nonumber
\end{equation}
Assuming that we interpret this measurement as experimental proof of lepton number violation, we would, as a next step, try to describe it by assuming the standard light neutrino exchange for the mediation of the decay~\cite{Bilenky:1987ty}. In this case, our expectation for the decay rate would be~\cite{Pascoli:2005zb}
\begin{equation}
\Gamma_\nu=G |\mathcal{M}^{0\nu}|^2 |m_{ee}|^2,
\label{eq:0nbb-rate_nu}
\end{equation}
where $G$ is a known phase space factor, $\mathcal{M}^{0\nu}$ is the nominal nuclear matrix element (NME), and $|m_{ee}|$ is the effective neutrino mass. The latter is a function of the smallest neutrino mass $m$, the mixing angles $\theta_{12}$ and $\theta_{13}$, the mass-squared differences $\mbox{$\Delta m^2_{21}$}$ and $\mbox{$|\Delta m^2_{31}|$}$, as well as the two Majorana phases $\alpha$ and $\beta$.\footnote{Note that $|m_{ee}|$ does not, contrary to some statements in the literature, depend on the Dirac CP-phase $\delta$, since this phase can be rotated away by redefining the third neutrino mass eigenstate $\nu_3$, see Refs.~\cite{Lindner:2005kr,Merle:2006du}.}
To determine the goodness of fit of the measured data with the standard neutrino-exchange mechanism, we will construct a $\chi^2$-function following the procedure outlined in Ref.~\cite{Pascoli:2005zb}, extended by information on $m_\beta^2$. In that paper, the authors have used the method of the covariance matrix given by
\begin{equation}
S_{ab}=\delta_{ab} \sigma_{a,\rm exp}^2 + \sum_i \frac{\partial T_a}{\partial x_i} \frac{\partial T_b}{\partial x_i} \sigma_i^2,
\label{eq:covariance}
\end{equation}
where $a,b\in \{ 1,2,3\}$, $\sigma_{a,\rm exp}$ are the experimental errors of the measured observables, $x_i\in \{ s_{12}^2, s_{13}^2, \mbox{$\Delta m^2_{21}$}, \mbox{$|\Delta m^2_{31}|$} \}$ are the relevant neutrino oscillation parameters, and $\sigma_i\in \{ \sigma(s_{12}^2), \sigma(s_{13}^2), \sigma(\mbox{$\Delta m^2_{21}$}), \sigma(\mbox{$|\Delta m^2_{31}|$}) \}$ are the corresponding standard deviations. Explicitly, the $T_a$'s are given by
\begin{equation}
T_1=\xi |m_{ee}|, \quad T_2=m_\beta^2, \quad T_3=\Sigma,
\label{eq:covariance_quantities}
\end{equation}
where
\begin{eqnarray}
|m_{ee}|&=&|m_1 c_{12}^2 c_{13}^2 + m_2 s_{12}^2 c_{13}^2 e^{2i\alpha} + m_3 s_{13}^2 e^{2i\beta}|,\nonumber\\
m_\beta^2&=&m_1^2 c_{12}^2 c_{13}^2 + m_2^2 s_{12}^2 c_{13}^2 + m_3^2 s_{13}^2,\label{eq:covariance_observables}\\
\Sigma&=&m_1+m_2+m_3.
\nonumber
\end{eqnarray}
The above observables are the main sources of information about the neutrino mass that we currently have:
\begin{itemize}
\item The effective neutrino mass $|m_{ee}|$ is, as just explained, measured in experiments that search for $0\nu\beta\beta$. We will take the GERDA experiment~\cite{Abt:2004yk} as an example.
\item The effective mass square of the electron neutrino, $m_\beta^2$, is measured in kinematical experiments on single $\beta$-decay, where KATRIN~\cite{Osipowicz:2001sq} is currently the most prominent upcoming example.
\item The effective sum $\Sigma$ of neutrino masses as measured in cosmological observations. This can be done through measurements of the cosmic microwave background (CMB) radiation, where we take the Planck satellite~\cite{Planck:2011ah} as example. However, the CMB measurements need to be combined with results from other observations such as high redshift galaxy surveys, baryon acoustic oscillations, or of Type Ia supernovae~\cite{Hannestad:2006zg,Hannestad:2007cp} to yield robust results.
\end{itemize}
Using these observables, the full $\chi^2$-function is given by
\begin{equation}
\chi^2(m,F, \alpha, \beta)=\min\limits_{\xi \in [1/\sqrt{F},\sqrt{F}]} v^T S^{-1} v,
\label{eq:chi2_total}
\end{equation}
where $v^T=(T_1-T_1^{\rm obs},T_2-T_2^{\rm obs},T_3-T_3^{\rm obs})$, and $T_a^{\rm obs}$ are the experimental values of the measured quantities. In particular, in the case of \mbox{$0\nu\beta\beta$}, we have
\begin{equation}
T_1^{\rm obs} = |m_{ee}|_{\rm obs}= \sqrt{\frac{\Gamma_{\rm obs}}{G}}\frac{1}{|\mathcal{M}^{0\nu}|}, \quad \sigma_{1, {\rm exp}} = \sigma (|m_{ee}|) = \frac12\frac{\sigma_{\rm obs}}{\Gamma_{\rm obs} }|m_{ee}|_{\rm obs}, \quad \xi=\frac{|\mathcal{M}^{0\nu}_{\rm true}|}{|\mathcal{M}^{0\nu}|}.
\label{eq:chi2_onbb_add}
\end{equation}
Here, $|\mathcal{M}^{0\nu}|$ is the nominal value of the NME used in the analysis to extract the experimental value of the effective mass, whereas $|\mathcal{M}^{0\nu}_{\rm true}|$ is its (unknown) true value. Furthermore, $F$ parametrizes the uncertainty in the NME, since it forces the $\chi^2$-minimization to treat all values of the NME between $|\mathcal{M}^{0\nu}|/\sqrt{F}$ and $|\mathcal{M}^{0\nu}|\sqrt{F}$ on the same footing, corresponding to a flat prior. As representative values we will choose $F=1$, $2$, and $3$, where $F=1$ corresponds to a perfect knowledge of the NME (cf.\ Ref.~\cite{Pascoli:2005zb}).
Note that, depending on which information is available, it might also be required to reduce Eq.~\eqref{eq:chi2_total} to incorporate, e.g., $|m_{ee}|$ and $\Sigma$ only, but not $m_\beta^2$. From Eq.~\eqref{eq:chi2_total}, it is possible to derive 1$\sigma$-ranges for the smallest neutrino mass $m$.
\section{\label{sec:scenarios}The benchmark scenarios}
\begin{figure}[t!]
\centering
\begin{tabular}[ht]{lr}
\includegraphics[width=0.47\textwidth]{GERDA_Reach.eps} &
\includegraphics[width=0.47\textwidth]{GERDA_Scenarios.eps}
\end{tabular}
\caption{\label{fig:scenarios} The experimental situation in a few years from now and the benchmark scenarios (A,B,C) to be considered.}
\end{figure}
The experimental situation in a few years from now is depicted in the left panel of Fig.~\ref{fig:scenarios}, which shows the range of the effective mass $|m_{ee}|$ in $0\nu\beta\beta$ as a function of the smallest neutrino mass $m$ for both orderings, normal ($m_1 < m_2 < m_3$) and inverted ($m_3 < m_1 < m_2$). The inner (outer) bands show the respective $1\sigma$ region for a value of $\sin^2 \theta_{13}=0$ (for the Double Chooz limit in case of a negative signal, $\sin^2 \theta_{13}<0.0076$~\cite{Palomares:2009wz}), while for the other oscillation parameters we have used the best-fit values and (symmetrized) standard deviations from Ref.~\cite{Schwetz:2011qt}. Obviously, the effect of a tiny but non-zero mixing angle $\theta_{13}$, is very small, in agreement with Refs.~\cite{Minakata:2001kj,Merle:2007hv}. In fact, for inverted ordering, the outer band (lighter yellow) is barely visible. The only effect of a non-zero $\theta_{13}$ shows up for the region of normal hierarchy (normal ordering with very small $m_1$), which cannot be probed in the near future. Accordingly, the dependence of the effective mass on the Majorana phase $\beta$ is very weak~\cite{Lindner:2005kr}.
We have also depicted the expected sensitivities of future experiments. First of all, the GERDA experiment will set different limits on the half-life of $0\nu\beta\beta$ in three different phases I, II, and III, which are given by $3\cdot 10^{25}$~years~\cite{Jochum:2010zz}, $1.5\cdot 10^{26}$~years~\cite{Jochum:2010zz}, and $2.8\cdot 10^{27}$~years~\cite{GERDA-Talk}. To translate these values into sensitivities on $|m_{ee}|$, one has, in principle, to know the correct NME. Since we do not want to enter a discussion about better and worse values of this quantity, we have taken, for simplicity, the most optimistic values from Ref.~\cite{Abt:2004yk} for phases I and II. As the reach of phase III is not completely clear yet (and hence, there is, to our knowledge, no official number from the GERDA collaboration), we have taken the benchmark value $|m_{ee}|=0.01$~eV, which is in agreement with recent estimates~\cite{GERDA-Talk}.
We have also depicted limits from complementary experiments that can yield information on the absolute neutrino mass scale. The only model-independent way to measure the neutrino mass is from simple kinematics, which will be performed by the KATRIN experiment~\cite{Osipowicz:2001sq}. Depending on the statistics used, one obtains different predictions for the sensitivity of the KATRIN experiment~\cite{Host:2007wh}, and we have used the slightly more optimistic value obtained by using Bayesian statistics, $m_\beta=0.17~{\rm eV}$ (which translates into just the same number for the smallest neutrino mass). The Planck experiment is supposed to probe the effective sum $\Sigma$ of light neutrino masses down to about $0.1$~eV~\cite{Hannestad:2006zg}, which we have taken as a benchmark value (and which translates into a limit on the smallest neutrino mass of roughly 0.02~eV).\footnote{Note that one might have to be careful when taking the cosmological observations at face value~\cite{Maneschg:2008sf}, since unknown systematic errors might be involved.}
Based on these numbers, we have chosen our three benchmark scenarios indicated by the solid green lines in the right panel of Fig.~\ref{fig:scenarios}, which in principle correspond to the three basic situations that are possible after the next generation of experiments:\footnote{However, note that there might of course be some subtleties involved, in particular in cases where the experimental results point to regions in parameter space which may correspond to more than one of our scenarios. We do not take into account such peculiarities, as long as there is no real measured signal which indeed falls in exactly such a region.}
\begin{enumerate}
\item[A:] {\it negative signal}\\
In this situation, no signal of \mbox{$0\nu\beta\beta$}\ has been measured by GERDA. No matter if we have additional information on the neutrino mass by kinematic measurements or by cosmology, we do not know if neutrinos are Dirac or Majorana particles. Correspondingly, we cannot use $|m_{ee}|$ to constrain $m_\beta$ and/or $\Sigma$, or vice versa. Concerning new physics, we can only give an upper bound on the size of the corresponding amplitude.
\item[B:] {\it consistent positive signal}\\
Here, GERDA has measured a certain rate of \mbox{$0\nu\beta\beta$}\, and the corresponding range obtained for $|m_{ee}|$ is consistent with complementary positive signals from KATRIN and Planck, as in the ``$\mathcal{QD}$ scenario'' of Ref.~\cite{Maneschg:2008sf}. In this case, one can fit all signals with the minimal scenario and we expect strong limits on contributions from further new physics. Depending on which term dominates, this situation will lead to a strong upper bound on $|\mbox{$\epsilon$}_i|$ or on the interference term $|\mbox{$\epsilon$}_i \cos \phi_i|$, cf.\ discussion in Sec.~\ref{sec:Analysis_B}.
\item[C:] {\it inconsistent positive signal}\\
Finally, we consider the case where GERDA has measured a non-zero rate of \mbox{$0\nu\beta\beta$} , which does, however, disagree with the values of $|m_{ee}|$ consistent with the measurements by KATRIN and Planck. In particular, these measurements are not consistent with a \mbox{$0\nu\beta\beta$}\ rate as high as measured. Then, we are forced to conclude that $0\nu\beta\beta$ is actually dominated by the contributions from new physics. In such a situation, one can estimate or even measure the magnitude of $|\epsilon_i|$.
\end{enumerate}
These three benchmark scenarios will be used in the following to derive the respective information on the contributions of point-like new physics to $0\nu\beta\beta$.
Finally, we want to remark that we could, of course, have chosen a different selection of experiments, in particular of future and upcoming searches for $0\nu\beta\beta$ (e.g.\ EXO~\cite{Danilov:2000pp,Conti:2003av}, Majorana~\cite{Gaitskell:2003zr}, or MOON~\cite{Ejiri:1999rk,Shima:2008zz}) or for the kinematical determination of the neutrino mass (e.g.\ MARE~\cite{Nucciotti:2010tx}). When having a look at the next five years, however, we consider our selection of experiments and scenarios to be both, representative and realistic.
\section{\label{sec:Rates}Half-lives and numerical analysis}
The full half-life obtained from the Lagrangian in Eq.~\eqref{eq:L_0nbb} in combination with the standard light neutrino exchange can be derived from Refs.~\cite{Pas:2000vn} and~\cite{Doi:1985dx}, and is given by
\begin{equation}
[T_{1/2}^{0\nu\beta\beta}]^{-1}=G_1 \left| \sum_{i=0}^3 \mathcal{\tilde M}_i \right|^2 + G_2 \left| \sum_{i=4}^5 \mathcal{\tilde M}_i \right|^2+ G_3\ {\rm Re}\left[ \left( \sum_{i=0}^3 \mathcal{\tilde M}_i \right) \left( \sum_{i=4}^5 \mathcal{\tilde M}_i \right)^* \right].
\label{eq:total_rate}
\end{equation}
The sum consists of three parts, namely the squared terms originating from the amplitudes labeled $(0,1,2,3)$, the squared terms originating from the amplitudes labeled $(4,5)$, and the interference terms.\footnote{Note that, depending on the helicities of the final state electrons, several of the interference terms in Eq.~\eqref{eq:total_rate} can be zero. This point will be discussed in a moment.} Although all operators have two electrons in the final state, the dependence of the nuclear/atomic physics parts on the final state energies are different~\cite{Doi:1985dx}, which is the reason for the appearance of the three different phase space factors $G_{1,2,3}$~\cite{Pas:2000vn}. The subscript ``0'' denotes the part coming from light neutrino exchange, while the other subscripts label the corresponding operators in Eq.~\eqref{eq:L_0nbb}. Accordingly, one has
\begin{equation}
\mathcal{\tilde M}_i= \left\{
\begin{matrix}
\mathcal{M}_0 & {\rm for}\ i=0,\\
\mbox{$\epsilon$}_i \mathcal{M}_i & {\rm for}\ i>0.
\end{matrix}
\right.
\label{eq:mod_amps}
\end{equation}
Here, the amplitude for the light neutrino exchange is given by the standard expression~\cite{Doi:1985dx,Bilenky:1987ty,Schmitz:1997ag,Simkovic:2010ka},
\begin{equation}
\mathcal{M}_0=\mathcal{M}^{0\nu} \frac{m_{ee}}{m_e}, \quad {\rm with}~~ \mathcal{M}^{0\nu}=-\frac{g_V^2}{g_A^2} \mathcal{M}_{F,\nu}+\mathcal{M}_{GT,\nu},
\label{eq:light_nu_amp}
\end{equation}
and all the other NMEs $\mathcal{M}_i$ as well as the phase space factors are specified in Ref.~\cite{Pas:2000vn}.
Note that there might be CP- or other phases involved, i.e., the coefficients $\mbox{$\epsilon$}_i$ might effectively be complex numbers (where any overall phase will cancel out, but a relative phase can remain). Hence, it is necessary to think about how to deal with possible interference terms. Sometimes in the literature, such phases are ignored by assuming CP-invariance~\cite{Ribeiro:2007ud}, which then results in relative signs only, which could be plus or minus~\cite{Schmitz:1997ag}, leading to minimal or maximal cancellation of the partial amplitudes. We will discuss this point in the following.
First of all, one should ask the question when interference terms can appear at all. The key point is that states of a different helicity will not interfere, as they have a quantum number that distinguishes them, a fact that is well known for, e.g., $\mu^- \mu^+ \to e^- e^+$ scattering~\cite{Peskin:1995ev}. But how is the situation in a \mbox{$0\nu\beta\beta$} -transition? As is usual, we consider only a $0^+ \to 0^+$ nuclear transition, which is by far the dominant contribution~\cite{Doi:1985dx}. Hence, no matter which chiralities were chosen in Eq.~\eqref{eq:hadron_operators}, the decay will always have {\it the same} initial state, namely a nucleus with zero spin and positive parity. Also, the nuclear final state will have zero spin. Furthermore, since the $s$-wave approximation is taken for the final state electrons~\cite{Pas:2000vn}, they will have a zero orbital angular momentum. In such a situation, the conservation of angular momentum forces the final electron spins to point in opposite directions. However, this does not constrain the helicities of the final state electrons, since the nucleus itself can obtain a non-zero 3-momentum, which allows all electron helicities to be produced.\footnote{In other words: The correct helicities are enforced by the electrons being emitted in certain directions. However, these directions cannot be detected in an experiment like GERDA.} Accordingly, even though we start with the same initial state for all transitions, we can have {\it different} final states, which will {\it not} interfere.
What we can learn from this discussion is that the decisive point is {\it only} the helicities of the final state {\it electrons}, cf.\ Eq.~\eqref{eq:electron_operators}. However, the final state quarks do not matter (exactly as the initial state quarks do not matter), since they are ``hidden'' inside the nucleus, and their helicities are never measured. Hence, even though the initial or final state helicities in Eq.~\eqref{eq:hadron_operators} might differ on the quark level, it does not matter for the actual decay, since the quarks are only intermediate states.
This leads us to a question: Which of the operators in Eq.~\eqref{eq:L_0nbb} can actually have the same final state electron helicities as the standard light neutrino exchange process? This is easy to answer: Using the symmetries involved, one can see that, in the standard case, the electron current is given by $j_R$ from Eq.~\eqref{eq:electron_operators}, due to the double weak vertex on the Majorana neutrino line~\cite{Bilenky:1987ty}. This leads to two left-handed electrons in the final state, as to be expected from the left-handedness of the weak interactions.\footnote{Actually, we are sloppily setting helicity and chirality equal, which is strictly speaking only justified in the highly relativistic limit. Nevertheless, the kinetic energy of both electrons for $^{76}$Ge will be $2.04$~MeV~\cite{Schmitz:1997ag}, making them relativistic, but not highly relativistic. Although this might lead to some error, we have to take into account that the largest error will actually come from the problematic determination of the NMEs involved.}
We can conclude: The only operators that can interfere with the standard process are the ones with the coefficients $\mbox{$\epsilon$}_1$, $\mbox{$\epsilon$}_2$, and $\mbox{$\epsilon$}_3$, and even they can only interfere if the current $j_R$ is involved. Accordingly, one can constrain the following operator coefficients without having to care about interferences: $\mbox{$\epsilon$}_{1}^{xyL}$, $\mbox{$\epsilon$}_{2}^{xyL}$, $\mbox{$\epsilon$}_{3}^{xyL}$, $\mbox{$\epsilon$}_{4}^{xyz}$, and $\mbox{$\epsilon$}_{5}^{xyz}$, where $x,y,z \in \{ L,R \}$.
Note that interference effects between different modes of $0\nu\beta\beta$ have also been discussed in Refs.~\cite{Faessler:2011qw,Faessler:2011rv}. They both agree with our findings in that the important condition for the appearance of interference effects is the equality in the chiral structures of the new physics contribution and the light neutrino exchange diagram, and hence the equality of the corresponding phase space factors. Furthermore, Ref.~\cite{Faessler:2011rv} points out that the introduction of new physics contributions could easily lead to relative CP-phases, as we have also mentioned.
Finally, we want to note that interferences will, if they appear at all, only play a role in scenario~B. The reason is simply that for scenarios~A and~C, it is perfectly enough to consider the contribution from the respective effective operator only.
These considerations allow us to write down expressions for the half-lives for all three benchmark scenarios:
\begin{enumerate}
\item[A:] {\it negative signal}\\
Here, only the contribution from the effective operators can be constrained. The expression we need is
\begin{equation}
[T_{1/2}^{0\nu\beta\beta}]^{-1}_{{\rm A},i}=\tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2,
\label{eq:T12_A}
\end{equation}
where $i=1,2,3,4,5$, $\tilde G_{1,2,3}=G_1$, and $\tilde G_{4,5}=G_2$.
\item[B:] {\it consistent positive signal}\\
In this case, we will also need to consider the contribution coming from the standard light neutrino exchange and obtain
\begin{eqnarray} \label{eq:T12_B}
[T_{1/2}^{0\nu\beta\beta}]^{-1}_{{\rm B},i} &=& G_1 \left| \mathcal{M}_0 \right|^2\ +\ \tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2 \\
& & \left( +\ 2 G_1 \cos \phi_i \cdot |\mbox{$\epsilon$}^{xyR}_i| \left| \mathcal{M}_0 \right| \left| \mathcal{M}_i \right|\ {\rm for}\ i=1,2,3\right), \notag
\end{eqnarray}
where $i=1,2,3,4,5$, and $\phi_i\equiv \phi^{xyR}_i$ is the relative phase between the two different contributions in the relevant cases. Of course, we cannot give an explicit expression for $\phi_i$ as long as the underlying theory is unknown, but e.g.\ in the case of the exchange of new heavy active neutrinos, this would correspond to a new Majorana phase $\gamma$ in addition to the two phases $\alpha$ and $\beta$ for light neutrinos.
\item[C:] {\it inconsistent positive signal}\\
Also here, we only need the contribution from the effective operators, which yields the same formula as for scenario A,
\begin{equation}
[T_{1/2}^{0\nu\beta\beta}]^{-1}_{{\rm C},i}=\tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2.
\label{eq:T12_C}
\end{equation}
\end{enumerate}
These scenarios will now be analyzed numerically.
\subsection{\label{sec:Analysis_A}Scenario A: Negative signal}
This scenario is the easiest one to analyze. The key point is that, in the case of a negative signal, we cannot use any external information on the neutrino parameters. The reason is that, even though we could in principle calculate a range for the effective mass $|m_{ee}|$ from data obtained by neutrino oscillation experiments in combination with kinetic and/or cosmological neutrino mass measurements, lepton number violation would not have been experimentally established, and hence, we would not even know if $|m_{ee}|$ has any physical meaning. Neutrinos could be Dirac particles, still in agreement with all experimental results to date.\footnote{Of course, this will change if we know from any other type of experiment that lepton number violation exists and that neutrinos are Majorana fermions. However, such a situation is not very likely to be present before data from the GERDA experiment will be available~\cite{Merle:2006du}, so we do not consider this case here.}
The only decent statement we can make in such a situation is that the contribution $\Gamma_\epsilon$ to \mbox{$0\nu\beta\beta$}\ from effective point-like operators can be at most as large as the upper bound on the decay rate measured in the experiment, i.e.,
\begin{equation}
\Gamma_\epsilon < \Gamma_{\rm bound}.
\label{eq:eps-range_negative}
\end{equation}
Note that this limit is a conservative one in the sense that we allow the contribution from the effective operators to be the dominant one, while any possible interference terms or further contributions are neglected. One could, of course, improve the bounds by assuming additional contributions rendering the point-like interactions much less dominant. However, the corresponding limit would only hold under this assumption and would hence be much more model-dependent.
Furthermore, Eq.~\eqref{eq:eps-range_negative} should be viewed as an estimate rather than as a strict bound at a certain confidence level, due to the intrinsic uncertainties associated with the computation of the NMEs. We will comment on this point in Sec.~\ref{sec:Caveat}.
In our analysis, we will always assume a single term from Eq.~\eqref{eq:L_0nbb} with a certain chirality structure to be dominant\footnote{This is what is called ``on-axis evaluation'' in Ref.~\cite{Pas:2000vn}.}, e.g.\ $\epsilon_5^{LRL} J_L^\mu J_R j_{L,\mu}$. This will lead to a certain decay rate $\Gamma_{\epsilon_5^{LRL}}$ that is a function of the corresponding coefficient $\epsilon_5^{LRL}$. Applying Eq.~\eqref{eq:eps-range_negative} will then yield an upper bound on $\epsilon_5^{LRL}$ .
A similar analysis has been done in Ref.~\cite{Pas:2000vn} for a 90~\%~C.L.\ limit obtained by the Heidelberg-Moscow experiment ($T_{1/2}^{0\nu\beta\beta}>1.8\cdot 10^{25}$~years, Refs.~\cite{Pas:2000vn,Baudis:1999xd}), which we have also reproduced as a cross check, using the same numbers as in the original reference.\footnote{Note that we could reproduce all numbers from Ref.~\cite{Pas:2000vn} except for the bound on $|\mbox{$\epsilon$}_3^{LLz,RRz}|$, where our value is smaller by a factor of two.}
To adjust the numbers to our scenario A, we have taken the 90~\%~C.L.\ limit for GERDA phase III in the case of a negative signal, $(T_{1/2}^{0\nu\beta\beta})_{\rm A}>2.8\cdot 10^{27}$~years~\cite{GERDA-Talk}. The nuclear radius for $^{76}$Ge has been estimated as $R \simeq 1.2 \sqrt[3]{76}$~fm $\simeq 5.1$~fm, and the numerical values for the phase space factors have been taken from Ref.~\cite{Pas:2000vn}. However, it was necessary to update the values of the NMEs, since there have been many developments within the last years, in particular in what regards the inclusions of short-range correlations and higher order nuclear currents~\cite{Fedor}. Taking the calculations from Refs.~\cite{Simkovic:2010ka,Fedor}, we can update the values from Ref.~\cite{Pas:2000vn} to be\footnote{Note that the matrix elements used in Ref.~\cite{Pas:2000vn} were based on the calculations from Ref.~\cite{Hirsch:1995ek}, up to a different sign convention in the definition of the Fermi part. The values from Refs.~\cite{Simkovic:2010ka,Fedor}, however, were based on the calculations from Ref.~\cite{Simkovic:1999re}, which uses a normalization for the individual Fermi and the Gamow--Teller parts that is smaller by a factor of $m_e/m_p$, but the full NME is of course of the same order of magnitude.}
\begin{equation}
\mathcal{M}_{F,N}=6.8\cdot 10^{-2} \quad {\rm and} \quad \mathcal{M}_{GT,N}=1.7\cdot 10^{-1}.
\label{eq:NME_updates}
\end{equation}
Note that we also use $g_A=1.25$, as suggested in Ref.~\cite{Simkovic:2010ka}. In addition, all other numerical values are taken from Ref.~\cite{Pas:2000vn}.
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|c|c|c|c|}\hline
A & $|\mbox{$\epsilon$}_1|$ & $|\mbox{$\epsilon$}_2|$ & $|\mbox{$\epsilon$}_3^{LLz,RRz}|$ & $|\mbox{$\epsilon$}_3^{LRz,RLz}|$ & $|\mbox{$\epsilon$}_4|$ & $|\mbox{$\epsilon$}_5|$\\ \hline \hline
Bound & $1.6\cdot 10^{-8}$ & $9.5\cdot 10^{-11}$ & $1.2\cdot 10^{-9}$ & $7.3\cdot 10^{-10}$ & $7.9\cdot 10^{-10}$ & $7.1\cdot 10^{-9}$\\ \hline
\end{tabular}
\caption{\label{tab:results_A} The resulting upper bounds on the magnitudes of the coefficients of the effective operators in scenario A, where $z=L$ or $R$.}
\end{table}
The resulting upper bounds on the magnitudes of the coefficients of the effective operators are given in Tab.~\ref{tab:results_A}, where we have indicated the chirality structures only for the cases where the bounds depend on them. As to be expected for scenario A, the bounds on the coefficients will be stronger than the bounds based on the Heidelberg-Moscow limit by about one order of magnitude, since phase III of GERDA would yield a limit on the decay rate that is about two orders of magnitude better, and since the squared absolute values of the coefficients enter the decay rate.
Finally, we would like to also derive a bound on the paremeter $\lambda'_{111}$ in our RPV-SUSY example. Using Eq.~\eqref{eq:eps_RPV}, as well as the bound on $\epsilon_1^{RRR}$ from Tab.~\ref{tab:results_A}, it is easy to derive the value $\lambda'_{111}<0.06$ for the sparticle masses assumed, which is a reasonable value as compared to other bounds on this parameter~\cite{Chemtob:2004xr}.
\subsection{\label{sec:Analysis_B}Scenario B: Consistent positive signal}
In case of a positive signal from GERDA, one first needs to check if all the available data from all experiments is consistent with the assumption that light neutrino exchange is the only mechanism behind $0\nu\beta\beta$. We will also investigate situations in which only {\it one} experiment (KATRIN or Planck) can yield complementary information on the neutrino mass, which would be important if one of the two ceases to produce reliable results.
The main technical problem is how to separate the point-like contributions resulting from the interactions in Eq.~\eqref{eq:L_0nbb} from the dominant contribution of light neutrino exchange, which is a long-range force~\cite{Pas:1997fx}. Note that, in principle, there might also be further long-range contributions, such as new heavy $W$-bosons in addition to the ones in the SM, in which case the neutrino line would still be of long range. However, we do not consider such contributions here, but instead focus on the point-like ones that can be fully expressed by effective field theory, in order to be as model-independent as possible.
To perform the analysis, we first need to assume a certain measured rate for scenario~B. Taking the smallest neutrino mass to be $0.3$~eV (cf.\ Sec.~\ref{sec:scenarios}), we roughly obtain (by varying the neutrino oscillation parameters and the Majorana phases) an effective mass $|m_{ee}|$ between 0.1~eV and 0.3~eV, which translates into a half-life between $2.8\cdot 10^{25}$~years and $2.6\cdot 10^{26}$~years for an NME of $4.0$. For definiteness, we assume a certain value within this range to be observed, let us say
\begin{equation}
\left(T_{1/2}^{0\nu\beta\beta}\right)_{\rm B}\simeq 5.0\cdot 10^{25}~{\rm years}.
\label{eq:TBas}
\end{equation}
This is equivalent to a decay rate of
\begin{equation}
\Gamma_{\rm obs}= \frac{\ln 2}{\left(T_{1/2}^{0\nu\beta\beta}\right)_{\rm B}}\simeq 1.4\cdot 10^{-26}/{\rm year}.
\label{eq:GammaBas}
\end{equation}
This measurement will, of course, have a certain error, the estimate of which can be found in Ref.~\cite{Abt:2004yk} to be
\begin{equation}
\frac{\sigma_{\rm obs}}{\Gamma_{\rm obs}}\simeq 23~\%,
\label{eq:GERDA_acc}
\end{equation}
which leads to an absolute error on the decay rate of $\sigma_{\rm obs}=0.32\cdot 10^{-26}/{\rm year}$. Finally, comparing Eqs.~\eqref{eq:0nbb-rate_nu}, \eqref{eq:total_rate}, and~\eqref{eq:light_nu_amp}, one observes that $G=G_1 \ln 2 / m_e^2$.
This enables the calculation of the parameters to be used in the $\chi^2$-function in Eq.~\eqref{eq:chi2_total}, namely, from Eq.~(\ref{eq:chi2_onbb_add}),
\begin{equation}
|m_{ee}|_{\rm obs} \simeq 0.23 ~{\rm eV}, \quad \sigma (|m_{ee}|)\simeq 0.026~{\rm eV},
\label{eq:sig0nbb}
\end{equation}
where a representative nominal value of $|\mathcal{M}^{0\nu}|=4.0$ has been used for the NME~\cite{Abt:2004yk,Simkovic:2010ka}.
Consistent values for the observables measured by KATRIN and Planck can be found in Ref.~\cite{Maneschg:2008sf} for a smallest neutrino mass of $0.3$~eV: $(m_\beta^2)_{\rm obs}=(0.30~{\rm eV})^2$ and $\Sigma_{\rm obs}=0.91$~eV, respectively. Realistic values for the corresponding errors are given by $\sigma(m_\beta^2)=0.025~{\rm eV}^2$~\cite{Osipowicz:2001sq,Host:2007wh} and $\sigma(\Sigma)=0.05$~eV~\cite{Hannestad:2007cp}. For definiteness, we express all neutrino masses by the lightest neutrino mass $m$ using the formulas that hold for an {\it inverted} mass ordering. However, as we are with $m=0.3$~eV in the quasi-degenerate region of the neutrino masses, it will anyway not make much of a difference which ordering is used~\cite{Lindner:2005kr}. Finally, the values and uncertainties of the neutrino oscillation parameters we use as input are given in Tab.~\ref{tab:oscillation_params}.
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|}\hline
Parameter & Best-fit value \& 1$\sigma$ range & Reference\\ \hline \hline
$s_{12}^2$ & $0.32\pm 0.016$ & \cite{Schwetz:2011qt}\\ \hline
$s_{13}^2$ & $0.00\pm 0.0076$ & \cite{Palomares:2009wz}\\ \hline
$\Delta m_{21}^2$ & $(7.6\pm 0.19)\cdot 10^{-5}~{\rm eV}^2$ & \cite{Schwetz:2011qt}\\ \hline
$|\Delta m_{31}^2|_{\rm nor.}$ & $(2.5\pm 0.09)\cdot 10^{-3}~{\rm eV}^2$ & \cite{Schwetz:2011qt}\\ \hline
$|\Delta m_{31}^2|_{\rm inv.}$ & $(2.3\pm 0.10)\cdot 10^{-3}~{\rm eV}^2$ & \cite{Schwetz:2011qt}\\ \hline
\end{tabular}
\caption{\label{tab:oscillation_params} The best-fit values and (symmetrized) 1$\sigma$ ranges for the neutrino oscillation parameters used in the analysis. Note that, in the inverted hierarchy case, $|m_{ee}|$ is actually proportional to $\sqrt{|\Delta m_{31}^2|_{\rm inv.}}$ and depends strongly on $\theta_{12}$, whereas there is only a very weak dependence on $\theta_{13}$~\cite{Lindner:2005kr}. Accordingly, the former two observables would require a precise experimental determination in order to maximize the use of our analysis.}
\end{table}
Now we are ready to investigate the $\chi^2$-function, Eq.~\eqref{eq:chi2_total}, which will only be a function of the absolute neutrino mass scale (or, equivalently, the smallest neutrino mass $m$), the two Majorana phases $\alpha$ and $\beta$ (where the dependence on the latter phase is rather weak), and the accuracy $F$ to which we know the NME. As already mentioned, we investigate nine cases, namely the one where both KATRIN and Planck yield a result as well as the two scenarios in which only one of these experiments yields a useful measurement, and each of these three cases will be investigated for perfect knowledge of the NME ($F=1$) as well as for different degrees of imperfect knowledge ($F=2,3$).
\begin{figure}[t]
\centering
\begin{tabular}{lr}
\includegraphics[width=0.43\textwidth]{Chi2B_G.eps} &
\includegraphics[width=0.43\textwidth]{Chi2B_Full.eps}\\
\includegraphics[width=0.43\textwidth]{Chi2B_K.eps} &
\includegraphics[width=0.43\textwidth]{Chi2B_P.eps}
\end{tabular}
\caption{\label{fig:chi2} The ranges for the smallest neutrino mass $m_3$ for different cases.}
\end{figure}
Following the procedure from Ref.~\cite{Pascoli:2005zb}, we minimize the $\chi^2$-function with respect to $\alpha$, $\beta$, and $\xi$. The results are the functions of $m$ shown in Fig.~\ref{fig:chi2}, where we have plotted the behavior of the $\chi^2$-function for the case where only GERDA yields a result (pessimistic), as well as for the cases where also KATRIN and/or Planck yield results. Note that a measurement of \mbox{$0\nu\beta\beta$}\ only (upper-left panel) does not yield an unambiguous result for the neutrino mass scale. This is natural because of the intrinsic uncertainties involved, i.e., bad knowledge of the NME involved can, in certain regions of the parameter space, be compensated by a variation of the Majorana phase $\alpha$ (and $\beta$, although this dependence is subdominant). Hence, the flat prior for the exact value of the NME corresponds to a perfectly flat region in the $\chi^2$-function.\footnote{Note that we display the full $\chi^2$-function, rather than only $\Delta \chi^2$, since our assumption of consistent results always allows for (at least) one perfect fit point.} This is different in the case when all experiments yield information (upper-right panel): Then, KATRIN as well as Planck will point to a certain neutrino mass scale, and hence, there will be no flat region left in the $\chi^2$-plot, which means that we actually have a measurement of the neutrino mass, and we can determine the corresponding best-fit point as well as the 1$\sigma$ region. The qualitative behavior does not change when we have only one experiment in addition to the \mbox{$0\nu\beta\beta$}-signal: If there is only the kinematical determination (lower-left panel), the $\chi^2$-function becomes broader, but it is nevertheless possible to derive a consistent range for the neutrino mass. The case of a cosmological observation being the only information in addition to \mbox{$0\nu\beta\beta$}\ (lower-right panel) looks practically identical to the case where all measurements fit together. This is simply a reflection of the small uncertainty assigned to the observation of the sum $\Sigma$ of neutrino masses. However, note that there could be systematic errors involved in the cosmological data that we are not aware of, which could even lead to wrong conclusions about the actual value of the neutrino mass~\cite{Maneschg:2008sf}, which is why this scenario might be a little too optimistic in case there is no additional information by KATRIN.
We have taken into account different values of the degree of knowledge of the NME ($F=1,2,3$) in all plots. However, for the 1$\sigma$ region, this only plays a role for the (too) pessimistic scenario of having a positive signal from \mbox{$0\nu\beta\beta$}\ only, even though the neutrino mass would be in a range where KATRIN and Planck could both determine it. We will not consider this case further.
Using the $\chi^2$-functions obtained, one can derive the best-fit values and 1$\sigma$ errors for $m$ in all cases, shown in Tab.~\ref{tab:massfit_B}. As to be expected, we always obtain our assumed value of $m=0.3$~eV as best-fit point, which is a reflection of the fact that we consider a situation in which all experiments yield consistent results. Note that, for the $1\sigma$-regions, the uncertainty in the NME does not play a major role, since we have enough complementary information from the kinematical and from the cosmological determination of the neutrino mass. Furthermore, note that, although it seems that the KATRIN experiment does not yield much more information compared to using the cosmological data alone, it nevertheless cannot be overemphasized that experiments based on kinematics are the only {\it model-independent} probe for the absolute neutrino mass scale that we have. Depending on the outcome of the cosmological observations, it might be necessary to dismiss the data on $\Sigma$ completely, in case that inconsistencies in our cosmological model would arise.
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|c|}\hline
$F$ & KATRIN \& Planck & KATRIN only & Planck only \\ \hline \hline
$1,2,3$ & $0.30^{+0.015}_{-0.016}$~eV & $0.30^{+0.040}_{-0.046}$~eV & $0.30^{+0.017}_{-0.017}$~eV\\ \hline
\end{tabular}
\caption{\label{tab:massfit_B} The reconstructed values and 1$\sigma$ ranges for the smallest neutrino mass.}
\end{table}
Now we are finally in the position to derive bounds on the different $|\epsilon_i|$'s. Let us shortly summarize what we have done: First of all, we started with Eq.~\eqref{eq:T12_B}, where we have neglected the interference terms, as discussed at the beginning of Sec.~\ref{sec:Rates},
\begin{equation}
[T_{1/2}^{0\nu\beta\beta}]^{-1}_{{\rm B},i}= G_1 \left| \mathcal{M}_0 \right|^2\ +\ \tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2.
\label{eq:T12_B_later}
\end{equation}
Next, we have assumed a certain observed rate, Eq.~\eqref{eq:GammaBas}, with an error given by $\sigma_{\rm obs}$ in Eq.~\eqref{eq:GERDA_acc}.
Our goal is to derive bounds on the $|\epsilon_i|$'s, so we should re-arrange Eq.~\eqref{eq:T12_B_later} to obtain
\begin{equation}
\tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2=[T_{1/2}^{0\nu\beta\beta}]^{-1}_{{\rm B},i}- G_1 \left| \mathcal{M}_0 \right|^2=\frac{\Gamma_{\rm obs}-\Gamma_\nu}{\ln 2}.
\label{eq:T12_B_rearranged}
\end{equation}
Since, as we have shown, the measurement of the neutrino mass scale is consistent with that of GERDA, the whole decay rate could be attributed to the standard mechanism of light neutrino exchange, causing the right-hand side of Eq.~\eqref{eq:T12_B_rearranged} to vanish. However, the effective operators could still contribute within the error of the measurement, giving the estimate
\begin{equation}
\tilde G_i |\mbox{$\epsilon$}_i|^2 \left| \mathcal{M}_i \right|^2 < \frac{\sigma_{\rm obs}}{\ln 2}.
\label{eq:T12_B_bound}
\end{equation}
Finally, Eq.~\eqref{eq:T12_B_bound} can be rewritten to yield a bound on $|\epsilon_i|$:
\begin{equation}
|\mbox{$\epsilon$}_i| < \frac{1}{\left| \mathcal{M}_i \right|} \sqrt{ \frac{\sigma_{\rm obs}}{\tilde G_i\ \ln 2} }.
\label{eq:epsi_B_bound}
\end{equation}
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|c|c|c|c|}\hline
B & $|\mbox{$\epsilon$}_1| \ (!)$ & $|\mbox{$\epsilon$}_2| \ (!)$ & $|\mbox{$\epsilon$}_3^{LLz,RRz}| \ (!)$ & $|\mbox{$\epsilon$}_3^{LRz,RLz}| \ (!) $ & $|\mbox{$\epsilon$}_4|$ & $|\mbox{$\epsilon$}_5|$\\ \hline \hline
Bound & $5.8\cdot 10^{-8} $ & $3.5\cdot 10^{-10}$ & $4.5\cdot 10^{-9}$ & $2.7\cdot 10^{-9}$ & $2.9\cdot 10^{-9}$ & $2.6\cdot 10^{-8}$\\ \hline
\end{tabular}
\caption{\label{tab:results_B} The resulting upper bounds on the magnitudes of the coefficients of the effective operators in scenario B, where $z=L$ or $R$. Here, ``$(!)$'' denotes that there is a possibility of interference if the current $j_R$ is involved in the operator.}
\end{table}
The numerical results can be found in Tab.~\ref{tab:results_B}. One observes that the constraints for this scenario appear to be slightly worse than the ones from scenario~A (cf.\ Sec.~\ref{sec:Analysis_A}). However, we can see in Fig.~\ref{fig:scenarios} that, for scenario~A, we would actually need a successful phase~III of GERDA, while a situation like scenario~B could even be reached within phase~I, the sensitivity of which is worse by a factor of roughly 100. Since the decisive quantity for the bounds is the error rather than the decay rate itself, cf.\ Eq.~\eqref{eq:epsi_B_bound}, the bounds obtained would be strong even for phase~I. However, this is only possible if there is no inconsistency in the experimental results, cf.\ Sec.~\ref{sec:Analysis_C}.
Using the example from Eq.~\eqref{eq:eps_RPV} again, we now arrive at $\lambda'_{111}<0.12$, which is slightly worse than the value from Sec.~\ref{sec:Analysis_A}, due to the difference on the respective bounds (cf.\ Tabs.~\ref{tab:results_A} and~\ref{tab:results_B}).
One last important remark is that some of our results, i.e., those marked with ``$(!)$'' in Tab.~\ref{tab:results_B}, could be altered by the presence of an interference term. However, as already discussed, these terms will only play a role when two conditions are simultaneously fulfilled:
\begin{enumerate}
\item the effective operator contains the electron current $j_R$ (otherwise there can be no interference in the highly relativistic limit), and
\item the relative phase $\phi$ between the standard operator and the new physics contribution is not close to $\pi/2$.
\end{enumerate}
The first condition is fulfilled for only a handful operators [cf.~Eqs.~\eqref{eq:L_0nbb} and~\eqref{eq:electron_operators}], and even then there is still the possibility of cancellation due to the value of the relative phase (second condition). Nevertheless, we also give the bounds for such a situation. In those cases where interferences might be relevant, we expect that, unless $\cos \phi_i $ is very small in Eq.~\eqref{eq:T12_B}, the last term will dominate over the second one. Then, once again assuming that the standard mechanism of light neutrino exchange is the main contribution, one can calculate an approximate value for $|\mathcal{M}_0|$ and then derive bounds on the combination of parameters $|\mbox{$\epsilon$}^{xyR}_i \cos \phi_i|$ as
\begin{equation}
|\mbox{$\epsilon$}^{xyR}_i \cos \phi_i| < \frac12 \sqrt{\frac{\sigma_{\rm obs}}{\Gamma_{\rm obs}}} B_i \simeq 0.24 B_i, \quad i=1,2,3,
\end{equation}
where $B_i$ is the left-hand side of Eq.~(\ref{eq:epsi_B_bound}), the values of which are given in Tab.~\ref{tab:results_B}. The resulting bounds on $|\mbox{$\epsilon$}^{xyR}_i \cos \phi_i|$ are shown in Tab.~\ref{tab:results_B_int}. These bounds are seemingly stronger by a factor of roughly four than the ones without interferences, but one has to keep in mind that the cosines involved may very well be of the order of 0.1, which would, in turn, weaken the bounds on the $|\mbox{$\epsilon$}^{xyR}_i|$'s.
Here, the RPV-SUSY parameter can only be constrained in connection with a CP-phase $\phi_\lambda$~\cite{Faessler:2011rv}. The corresponding bound from Tab.~\ref{tab:results_B_int} translates into $|\lambda'_{111} \cos \phi_\lambda|<0.028$, which could turn out to be particularly strong in case the value of the CP-phase is predicted in a certain model.
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|c|c|}\hline
B (interf.) & $|\mbox{$\epsilon$}^{xyR}_1 \cos \phi_1|$ & $|\mbox{$\epsilon$}^{xyR}_2 \cos \phi_2|$ & $|\mbox{$\epsilon$}^{LLR,RRR}_3 \cos \phi_3|$ & $|\mbox{$\epsilon$}^{LRR,RLR}_3 \cos \phi_3| $ \\ \hline \hline
Bound & $1.4\cdot 10^{-8} $ & $8.4\cdot 10^{-11}$ & $1.1\cdot 10^{-9}$ & $6.5\cdot 10^{-10}$ \\ \hline
\end{tabular}
\caption{\label{tab:results_B_int} The resulting upper bounds on the combination of parameters $|\mbox{$\epsilon$}^{xyR}_i \cos \phi_i|$ in scenario B, assuming that the interference terms dominate. Here, $x,y=L$ or $R$.}
\end{table}
\subsection{\label{sec:Analysis_C}Scenario C: Inconsistent positive signal}
The final case to consider is a positive signal in GERDA whose rate cannot be described by neutrino physics only. As an example, we take a measured decay rate of $( T_{1/2}^{0\nu\beta\beta} )_{\rm C}=1.26\cdot 10^{27}$~years (well below the GERDA phase II limit), which would translate into an effective mass of $|m_{ee}|=0.045$~eV for an NME of 4.0. In case of a negative signal in KATRIN and Planck [and under the implicit assumption that there are no unknowns in cosmology which would affect $\Sigma$ in a way that it is not given by the expression in Eq.~\eqref{eq:covariance_observables} anymore], we would practically exclude inverted mass ordering, since the minimum value of $\Sigma$ in this case is $\Sigma_{\rm min}^{\rm IH}=0.1$~eV, which is just excluded by a negative signal from cosmology.
In this case, the contribution of the light neutrino exchange to the decay rate would be smaller than the contribution from the effective operators by some factor. Then, under the assumption that the neutrino contribution can be completely neglected, one can actually derive a value (and, in principle, allowed ranges), which serves as estimate for the coefficients $\mbox{$\epsilon$}_i$. Note that this would be in particular valid in case that the parameters are such that $|m_{ee}|$ is practically zero~\cite{Lindner:2005kr}. In Fig.~\ref{fig:chi2_C}, the minimized $\chi^2$-function as a function of the smallest neutrino mass is shown for normal and inverted mass ordering. We can see that, although we could in principle explain such a high \mbox{$0\nu\beta\beta$}\ rate for a certain range of the neutrino mass, this range does not fit with the additional information obtained by Planck which strongly disfavors the mass ranges required by GERDA. Accordingly, the minima of the full $\chi^2$-functions will be at least around 4, indicating that the results are inconsistent.
\begin{figure}[t]
\centering
\begin{tabular}{lr}
\includegraphics[width=0.43\textwidth]{Chi2C_All_IH.eps} &
\includegraphics[width=0.43\textwidth]{Chi2C_GP_IH.eps} \\
\includegraphics[width=0.43\textwidth]{Chi2C_All_NH.eps} &
\includegraphics[width=0.43\textwidth]{Chi2C_GP_NH.eps}
\end{tabular}
\caption{\label{fig:chi2_C} The minimized $\chi^2$-function as a function of the smallest neutrino mass, which is $m_1$ for normal mass ordering and $m_3$ for inverted mass ordering.}
\end{figure}
We will now use Eq.~\eqref{eq:T12_C} to estimate the magnitudes of the $\mbox{$\epsilon$}_i$'s. The \mbox{$0\nu\beta\beta$}\ decay rate $\Gamma_{\rm obs}$, cf.\ Eq.~\eqref{eq:GammaBas}, will translate into estimates of the sizes of the $\mbox{$\epsilon$}_i$'s,
\begin{equation}
|\mbox{$\epsilon$}_i |_{\rm obs} \simeq \frac{1}{\left| \mathcal{M}_i \right|} \sqrt{ \frac{\Gamma_{\rm obs}}{\tilde G_i\ \ln 2} }.
\label{eq:eps_acc}
\end{equation}
Using the same formulas as in Sec.~\ref{sec:Analysis_A}, we obtain estimates in the case of scenario~C shown in Tab.~\ref{tab:results_C}.
\begin{table}[t]
\centering
\begin{tabular}{|c||c|c|c|c|c|c|}\hline
C & $|\mbox{$\epsilon$}_1|$ & $|\mbox{$\epsilon$}_2|$ & $|\mbox{$\epsilon$}_3^{LLz,RRz}|$ & $|\mbox{$\epsilon$}_3^{LRz,RLz}|$ & $|\mbox{$\epsilon$}_4|$ & $|\mbox{$\epsilon$}_5|$\\ \hline \hline
Estimate & $2.3\cdot 10^{-8}$ & $1.4\cdot 10^{-10}$ & $1.8\cdot 10^{-9}$ & $1.1\cdot 10^{-9}$ & $1.2\cdot 10^{-9}$ & $1.1\cdot 10^{-8}$\\ \hline
\end{tabular}
\caption{\label{tab:results_C} The resulting estimates for the magnitudes of the coefficients of the effective operators in scenario C, where $z=L$ or $R$.}
\end{table}
We conclude that one could, in principle, constrain the coefficients of the effective operators very well by simply assuming that the contributions from neutrino physics are completely subdominant. As to be expected, values slightly smaller than the bounds obtained in Ref.~\cite{Pas:2000vn} could be measured.
Accordingly, one can also derive the magnitude of the RPV-SUSY parameter in this scenario, which turns out from Tab.~\ref{tab:results_C} to be $\lambda'_{111}\simeq 0.074$. In principle, this method could be used to measure the amount of $R$-parity violation present.
\subsection{\label{sec:Caveat}On the importance of the knowledge of the nuclear physics}
There is an important caveat in our argumentation: In the above analyses, we have (in connection with the effective operators) always used the values for the NMEs derived from Refs.~\cite{Simkovic:2010ka,Fedor} (which update the ones given in Ref.~\cite{Pas:2000vn}). However, in reality these values will be uncertain. Different working groups might obtain different numbers, which could then be compared to the experimental results. A more advanced analysis that takes this into account has been performed in our scenarios~B and~C for the light neutrino exchange diagram. However, for the effective operators we do not want to guess some values of the NMEs that may be more or less likely, as long as no further calculations exist. One still has to keep in mind that exactly these computations might become very important in case that, after the next generation of experiments will have been completed, we would indeed end up in a situation that is similar to one of our scenarios.
The goal of this paper was to demonstrate how strong the bounds on point-like new physics can become in the case of a positive or negative future signal of \mbox{$0\nu\beta\beta$}. This has been achieved by the estimates obtained in Secs.~\ref{sec:Analysis_A}, \ref{sec:Analysis_B}, and~\ref{sec:Analysis_C}. However, this does not cure the problem of the relatively poor knowledge of the nuclear physics involved. Although this might not matter too much in a phenomenological analysis like the one presented here, it will be crucial for the true evaluation of future data -- a point that cannot be overemphasized. Hence, our results can very well serve as a guideline for how existing bounds on the effective operators will improve for different realistic situations. However, the numbers obtained suffer from the uncertainties involved in the computations of the nuclear matrix elements and can, accordingly, not be treated as strict statistical bounds.
\section{\label{sec:Conc}Conclusions}
We have investigated point-like operators contributing to neutrino-less double beta decay using the formalism of effective field theory. In order to focus on situations that can be tested in the near future, we have defined three scenarios, A, B, and C, which correspond to a negative double beta signal, a positive double beta signal that is consistent with complementary experimental results, and a positive double beta signal that is inconsistent with complementary experimental results, respectively. For each scenario, we have determined the bounds on (or estimates for) the strengths $\mbox{$\epsilon$}_i$ of each point-like operator that could contribute to the decay. Our scenario A is merely an update of bounds that have already been present in the literature, while our scenarios B and C have, to our knowledge, not been investigated before in this context. In case neutrino-less double beta decay is not detected at all, the current bounds on the $\mbox{$\epsilon$}_i$'s can be improved by roughly one order of magnitude. The most interesting case arises if a positive signal is consistent with all complementary results: Then, the possible room for further new physics contributions to double beta decay on top of the standard light neutrino exchange is essentially determined by the error in the measured rate, which is much smaller than the rate itself. In case of an inconsistent positive signal, one can even give estimates of the coefficients $\mbox{$\epsilon$}_i$. Our results do, however, suffer from the lack of knowledge of the underlying nuclear physics, which is a well-known problem in studies of neutrino-less double beta decay. Accordingly, the bounds that we have derived will be made much more robust once the nuclear matrix elements are known to a better precision. In that case, our bounds would depend directly on the experimental accuracy and could be significantly improved by future attempts to measure the rate of neutrino-less double beta decay more precisely.
\section*{\label{sec:Ack}\noindent Acknowledgments }
We would like to thank M.~D\"urr for useful discussions, and we are especially grateful to F.~\v{S}imkovic for kindly providing us with his values obtained for the partial NMEs. This work has been supported by the Swedish Research Council (Vetenskapsr{\aa}det), contract no.\ 621-2008-4210 (T.O.) and by the Royal Institute of Technology (KTH), project no.\ SII-56510 (A.M.).
\bibliographystyle{./apsrev}
|
\section{\label{sec:level1}Introduction}
There is currently a lot of experimental and theoretical interest in rare-earth
pyrochlore magnets of stoichiometry, R$_2$M$_2$O$_7$, where R is a rare-earth
atom and M is an atom that has an available M$^{4+}$ state, such as tin or
titanium. Much of the interest has focused on `spin-ice'\cite{7}, the
Ho$_2$Ti$_2$O$_7$ and Dy$_2$Ti$_2$O$_7$ compounds, due to the proposal that
they can be considered to be analogue `monopole' systems\cite{8}. The
original motivation for studying these systems was the macroscopic degeneracy
expected for the isotropic Heisenberg model\cite{9}. The compounds
Gd$_2$Ti$_2$O$_7$\cite{10} and Gd$_2$Sn$_2$O$_7$\cite{11} are systems that
are expected to conform to this model at intermediate temperature. The issue
of how this residual degeneracy is lifted at very low temperature then
surfaced and further work understanding the dipolar interaction was then
relevant\cite{12}. Spin-ice is a system with a strong crystal-field
interaction forcing the spins to point along the natural crystallographic
directions whereas the gadolinium compounds have essentially no anticipated
crystal field (subject to some unexpected ESR issues\cite{13}). To complete
the set it would be nice to have a material where the local crystal-field
promoted spins perpendicular to the natural crystallographic directions, and
the compound Er$_2$Ti$_2$O$_7$ was thought to play this role\cite{14}.
Theoretical crystal-field calculations\cite{5}, neutron scattering
experiments\cite{1,2,3} and specific heat measurements\cite{6} were all
brought forward to promote this picture. Indeed, a subtle `order from
disorder' calculation\cite{15} appeared to predict the observed experiments.
In this article we offer an alternative view that the spins are oriented quite
close to the natural crystallographic axes, but are actually slightly
distorted away into a point-symmetry broken orientation. This loss of
point-symmetry then leads to a continuous degeneracy which explains the
anomalous entropy still present at low temperatures.
Our proposal for the physics of Er$_2$Ti$_2$O$_7$ is essentially equivalent
to previous studies of multiple-{\bf q} magnetism in face-centre-cubic
magnets\cite{4}, which contain most of the fundamental ideas and concepts.
We organise the paper into three main sections; firstly a discussion of the
main experiments and how they lead us to our proposed ground-state, secondly
a theoretical model that can be used to understand the majority of the
experiments and thirdly a semi-classical spinwave analysis of our model. Our
model is clearly only approximately solved which causes clear `gaps' in our
reasoning. Finally we conclude.
\section{\label{sec:level2}Experiments and interpretation}
We commence with the specific heat measurements\cite{6}, which have been
repeated as a function of field\cite{2,3,16}, although the original
measurements are all that we require. These specific heat measurements lead
directly to three crucial physical facts. Firstly there is a
clear phase transition near in energy to the natural dipolar
energy-scale\cite{14}. Secondly there are only two states per atom involved in
this low temperature phase transition, a Kramer's doublet, and so we need only
deal with a pseudo-spin half. Thirdly the relevant low temperature magnetic
contribution is gapless and the observed power-law is consistent with a magnon
dispersion that goes linearly to zero at isolated points in reciprocal-space.
It is this third point which is totally unexpected: spin-orbit interactions
provide a coupling between the spin direction and the lattice which almost
invariably lead to a spinwave gap. In transition metals the spin-orbit
interaction is weak in comparison to the exchange\cite{17} and so the gap is
small, but in rare-earth magnets this interaction is usually huge and any
associated low energy mode is absent. The only likely failure for this
argument is when the atom has a pure isotropic spin, gadolinium for example, where this
anisotropy might be expected to be absent. Erbium would be expected to have a
huge anisotropy gap. Sometimes low energy modes are seen (although not
vanishing), in CeAs for example\cite{18}, however the interpretation for these
modes is different\cite{4}. These special modes occur in multiple-{\bf q}
magnetism, only found in frustrated systems whose magnetism breaks the
point-group symmetry of the underlying crystal. The modes are associated with
transferal of spin-density between distinct Bragg spots and correspond to local
spin rotations which are longitudinal antiferromagnetic in character. We are
proposing that it is one of these special modes that has gone soft in
Er$_2$Ti$_2$O$_7$.
The most important experiments for our development are those involving
neutrons. There is a powder diffraction experiment\cite{1}, a single-crystal
diffraction experiment in a restricted plane but with critical inelastic
measurements\cite{2}, a fairly complete single-crystal diffraction
experiment\cite{3} and a polarised neutron study\cite{19}. We will deal with
the elastic scattering first and then consider the inelastic separately.
Elementary investigation of neutron scattering data involves three critical
elements: the structure-factor, the form-factor and the orientational-factor.
The structure-factor variation stems from the different moment directions
occurring on distinct atoms in the magnetic unit cell, and for us is fairly
definitive. The form-factor arises from the orbital spread of the magnetic
moment about the site of the nucleus, and is both a complication and a
concern. The magnetic erbium core is in a very complicated state, dominated by
Hund's rules, it has a maximal total angular momentum which is parallel to its
maximal spin to provide $J=15/2$. The sixteen states are crystal field split
to a relevant pseudo-spin doublet. The additional crystal-field doublets are
visible using inelastic neutron scattering\cite{1}, but are at high energies
in comparison to the inate energy of the phase transition. Indeed, the
specific-heat measurements yield almost precisely this pseudo-spin degeneracy
before the higher energy states impinge\cite{6}. These
complex low energy states have a shape that will distort the
form-factor\cite{20}. We will ignore this distortion and presume that the
form-factor is spherically symmetric. This issue is usually not a worry,
because domain averaging reinstates the point-symmetry, however our proposed
magnetism breaks that point-symmetry and would highlight these effects. The
orientational-factor (that stems from the fact that neutrons can only scatter
from moments perpendicular to the direction of momentum transfer) is very
physically informative and marks the start of our analysis.
In simple terms, for any set
of spots with the same structure-factor, one looks for the smallest
(form-factor compensated) spot. The magnetic moment associated with that
structure-factor is then essentially parallel to the reciprocal-lattice-vector,
{\bf G}, of that spot. The neutron-scattering data then provides the
following information. The closest spots to the origin, $(111)$, $(1{\bar 1}{\bar
1})$, $({\bar 1}1{\bar 1})$ $({\bar 1}{\bar 1}1)$, are all quite small. The
second-closest spots to the origin, (200), (020) and (002) are all very small,
while the third-closest spots, (022), (202) and (220), are dominant. The
second observation is crucial, since it tells us that the three independent
components of spin-density that characterise the magnetism are parallel to the
Cartesian directions. The states are described mathematically by
\begin{eqnarray}
\label{mag}
{\bf m}({\bf R})=\mid {\bf m}\mid \big[ e^{{\bf \hat x}.{\bf R}2\pi i}\sin
\theta \cos \phi {\bf \hat x}\; \; \; \; \; \; \; \; \; \; \; \; &&\nonumber \\
\; \; \; \; \; \; \; \; \; \; \; \; +e^{{\bf \hat y}.{\bf R}2\pi i}\sin \theta
\sin \phi {\bf \hat y}+e^{{\bf \hat z}.{\bf R}2\pi i}\cos \theta {\bf \hat z}
\big] &&
\end{eqnarray}
where ${\bf R}$ is the position of the atom in natural units (corresponding to
the underlying face-centre-cubic Bravais lattice) and
pictorially by Fig.\ref{fig:1}. To limit to the depicted magnetism, we also
\begin{figure}
\includegraphics[height=7.2 cm, width=8.4 cm]{ertif1}
\caption{\label{fig:1} The spin-states consistent with vanishing Bragg spots
(200), (020) and (002).}
\end{figure}
require that there is no ferromagnetism in the system, which is controlled by
the almost vanishing (222) Bragg spot. Technically, these states are
equivalent to Type I face-centre-cubic antiferromagnetism\cite{4}, but
interestingly, the pyrochlore magnetism is easier to deduce than that of
face-centre-cubic. The key observation is that the $(111)$ spots provide
complementary information: this particular spot provides a phase which splits
the lattice up into planes perpendicular to the $(111)$ direction, each of
which has uniform phase. These planes happen to be alternating Kagom\'e and
sparse triangular planes with alternating signs. The key observation is that
these triangular planes make up one of the four sublattices, while the
Kagom\'e planes make up the other three. If the system is an antiferromagnet,
then the structure factor for this Bragg spot should be proportional to the
spin direction on the triangular sublattice. The experimental observation
that the $(111)$ Bragg spot is quite small then provides the information that
the spin on site zero of Fig.\ref{fig:1} is almost parallel to the $(111)$
direction. This clearly contradicts the endemic idea that the spins are
perpendicular to the natural crystallographic
directions\cite{1,2,3,5,14,15,16,19}.
The neutron experiments also clearly exhibit distinct magnetic
domains\cite{1,2,3}. This is clearly inconsistent with the idea that the
state is a pure triple-{\bf q} state (spins pointing towards the natural
crystallographic directions, defined by $\cos \Theta =\frac{1}{\surd 3}$),
because this state preserves the cubic symmetry. The magnetic field
experiments provide us with an essentially unique state. The experimental
evidence was initially conflicting; the first measurement\cite{1} tried to
imply domains from a jump in intensity for a single Bragg spot when a field was
applied. Domains correspond to {\it redistribution} of Bragg intensity, and so
if one peak grows then another should reduce, whereas a change in a single
Bragg spot should signal a fundamental change in magnetism and {\it not} a
change in domain structure. This experimental result was therefore anomalous
until the single-crystal measurement\cite{2} indicated an unexpected and
unexplained diffuse peak that enlarged the Bragg spot with the inclusion of the
field. This single-crystal work\cite{2} does show clear indications of
domains, but because not all of the magnetic Bragg scattering was observed it
is not unambiguous. The final fairly complete measurement\cite{3} measures all
three crucial peaks independently and they are seen to separate in an intensity
preserving way once the field is applied. We can fit the three intensities to
our Eq.(\ref{mag}) and we find that
\begin{equation}
\tan \phi =1\; \; \; \; \; \; \; \; \tan \theta \sim \frac{4}{3\surd 3}
\end{equation}
although no crystallographic significance should be attributed to our
approximation. This direction is quite close to $\frac{1}{\surd 6}(1,1,2)$,
which also holds no special significance, unlike the previously proposed,
$\frac{1}{\surd 6}(1,1,\bar 2)$, which is perpendicular to the natural
crystallographic directions. We have also analysed the $(111)$ spot\cite{2}
which appears consistent with this choice, subject to our crude choice of
form-factor, a pure hydrogenic $f$-state. The state we propose has tetragonal
magnetic symmetry and lies on the path that smoothly connects the pure
triple-{\bf q} state to the single-{\bf q} state\cite{4}.
Finally, we consider the inelastic neutron scattering\cite{2} together with an
ESR experiment\cite{16}. The low-energy gapless mode is clearly visible close
to the magnetic Bragg spots, and a second almost flat mode is also visible.
Perhaps the most intriguing observation is that when the field is applied
these magnons disappear, but reappear close to other point-group related Bragg
spots\cite{2}, a subtle domain issue that we do not currently try to explain.
Note that the position of this Goldstone mode is exactly equivalent to that
in single-{\bf q} CeAs\cite{18} or indeed in USb\cite{21} or UO$_2$\cite{22},
both of which are thought to be pure triple-{\bf q} states. In parallel to
this, the ESR experiments\cite{16} see analogous modes but chase them as a
function of field. The physical surprise here is that it is the flat mode that
softens and controls the loss of antiferromagnetism and not the low-energy
Goldstone-mode as might na\"ively be presumed. This is further evidence that
the low energy mode is not of the usual type.
\section{Modelling}
Modelling rare-earth magnetism properly is complicated. The dominance of
Hund's rules makes the orbital nature of the states complex and the local
crystal-field interactions destroy a simple isotropic-spin picture. In
previous work the case of Gd$_2$Ti$_2$O$_7$\cite{12} was considered, here we
anticipate a pure spin with no orbital complications and natural models. In
this section we develop a phenomenological model to try to describe the
observed behaviour of the erbium compound, and the natural models are severely
renormalised.
One major theoretical complication is the crystal-field interaction, which
projects the original large-J states onto a pseudo-spin half representation
that controls the low temperature phase transition. This projection strongly
renormalises the interactions and hence corrupts intuition. One can provide
some understanding using an elementary model for this projection. For a
half-integer J we can easily show that the projection onto the states
$\mid J,\pm J\rangle $ along the natural crystallographic directions, defined
by,
\begin{eqnarray}
{\bf \hat z}_0\equiv \frac{{\bf \hat x}+{\bf \hat y}+{\bf \hat z}}{\surd 3}
\; \; \; \; \; \; \; \; \; \;
{\bf \hat z}_1\equiv \frac{{\bf \hat x}-{\bf \hat y}-{\bf \hat z}}{\surd 3}
\; \; &&\nonumber \\
{\bf \hat z}_2\equiv \frac{-{\bf \hat x}+{\bf \hat y}-{\bf \hat z}}{\surd 3}
\; \; \; \; \; \; \; \;
{\bf \hat z}_3\equiv \frac{-{\bf \hat x}-{\bf \hat y}+{\bf \hat z}}{\surd 3}&&
\end{eqnarray}
yields the mapping
\begin{equation}
\label{para}
{\bf \hat J}_\alpha \mapsto 2J{\bf \hat z}_\alpha \left( {\bf \hat z}_\alpha .
{\bf \hat S}_\alpha \right)
\end{equation}
where ${\bf \hat J}_\alpha $ are the original angular-momentum operators and
${\bf \hat S}_\alpha $ are the pseudo-spin half operators. We can then
project any bilinear Hamiltonian onto an effective spin-half model. For the
case of Er$_2$Ti$_2$O$_7$ the moment along the crystallographic axes is thought
to be small\cite{5} and in addition the ordered moment is indeed observed to be small\cite{1}.
To analyse this system it therefore seems more natural to
project onto the states $\mid J,\pm \frac{1}{2}\rangle $ using the following mapping:
\begin{equation}
\label{perp}
{\bf \hat J}_\alpha \mapsto \left( J+\frac{1}{2}\right) {\bf \hat S}_\alpha
-\left( J-\frac{1}{2}\right) {\bf \hat z}_\alpha \left( {\bf \hat z}_\alpha .
{\bf \hat S}_\alpha \right)
\end{equation}
The role of the crystallographic axes is dominant in the first case,
Eq.(\ref{para}), and weakened in the second, Eq.(\ref{perp}). At
bilinear order, our projection onto pseudo-spin is quite general, if the two
coefficients in Eq.(\ref{perp}) are made parameters, depending only on the
idea that the crystal-field is dominated by a single natural direction.
We develop five natural interactions in our modelling. Four stem solely from
the crystallographic directions and bilinear coupling. The first, $H_0$, has
no pseudo-spin analogue, and the others, $H_1$, $H_2$ and $H_3$, provide the
dominant spin interactions. The fifth interaction, $H_4$, couples the spin
to the lattice and is crucial to our modelling. We elect to derive these
interactions by looking at the natural energies in our system.
The first energy we consider is the dipolar energy, this is not the strongest
interaction as both exchange and residual crystal-field are expected to
dominate, but it is very instructive. Working on a single tetrahedron, in the
presence of dominant antiferromagnetic Heisenberg interactions, the local
dipolar interaction can be shown to take the form of a linear combination of
two contributions
\begin{eqnarray}
H_4=\frac{1}{2}\Big[ \left( \hat S_0^x+\hat S_1^x-\hat S_2^x-\hat S_3^x\right)
^2\; \; \; &&\nonumber \\+\left( \hat S_0^y-\hat S_1^y+\hat S_2^y-\hat S_3^y
\right) ^2\; \; \; &&\nonumber \\+\left( \hat S_0^z-\hat S_1^z-\hat S_2^z+\hat
S_3^z\right) ^2\Big] &&
\end{eqnarray}
and
\begin{eqnarray}
H_0=\frac{1}{6}\Big[ \left( \hat S_0^x+\hat S_0^y+\hat S_0^z\right) ^2
+\left( \hat S_1^x-\hat S_1^y-\hat S_1^z\right) ^2\; \; \; \; \; \; \; \; &&
\nonumber \\ \; \; \; \; \; \; +\left( -\hat S_2^x+\hat S_2^y-\hat S_2^z\right)
^2+\left( -\hat S_3^x-\hat S_3^y+\hat S_3^z\right) ^2\Big] &&\nonumber \\
\equiv \frac{1}{2}\sum _\alpha \left( {\bf \hat z}_\alpha .{\bf \hat S}_\alpha
\right) ^2\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
where the coefficient of $H_0$ is six times that of $H_4$ for the local
interaction. Using Madelung techniques\cite{12}, it can also be shown that
the dipole interaction, restricted to the ${\bf q}$=${\bf 0}$ subspace, also
takes the same form (for the {\it static} issues) but with $H_0$ slightly
more important. We also elect to incorporate the interaction
\begin{equation}
H_3=\frac{1}{8}\left( \sum _\alpha {\bf \hat z}_\alpha .{\bf \hat S}_\alpha
\right) ^2
\end{equation}
which is generated by our projections. Treating these Hamiltonians as though
they were initially functions of ${\bf \hat J}_\alpha $, applying the
maximal spin projector then yields
\begin{equation}
H_4\mapsto 16J^2H_3\; \; \; \; \; H_0\mapsto 4J^2H_0\; \; \; \; \; H_3\mapsto
4J^2H_3
\end{equation}
All that remains for this case is $H_0$ and $H_3$. Although the Hamiltonian
$H_0$ is crucial for the original angular momentum, once we arrive at our
spin-half pseudo-spin, this interaction reduces to a constant and can be
ignored. As we only look at semi-classical solutions to our Hamiltonians, we
are forced to keep $H_0$ active, although technically we need only the single
Hamiltonian, $H_3$, for this limit. Although $H_0$ is lost, there is still
effectively a residual crystal-field interaction, $H_3$, promoting the natural
crystallographic directions, it is just no longer local. Spin-ice\cite{7}
is thought to be well represented by the maximal projected pseudo-spin. The
spin component parallel to the crystallographic direction, ${\bf \hat z}_\alpha
.{\bf \hat S}_j$, is then a conserved quantity and the pseudo-spin always
points in these directions. The residual interaction energy on a tetrahedron
is described by $H_3$ and is minimised by a `two-in and two-out'
configuration.
We will now consider the minimal spin-projector model. It is slightly more
complicated to analyse but provides
\begin{eqnarray}
\label{ten}
H_4\mapsto \left(J(J+1)+\frac{1}{4}\right) H_4-4\left( J(J+1)-\frac{3}{4}
\right) H_3&&\nonumber \\H_0\mapsto H_0\; \; \; \; \; \; \; \; \; \; \; \; \;
H_3\mapsto H_3\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
The $H_3$ contribution now appears with the opposite sign. We observe that in the
large J limit Eq.(\ref{ten}) becomes
\begin{eqnarray}
H_4-4H_3=\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \;
\; \; \; \; \; \; \; &&\nonumber \\
\frac{1}{12}\Big[ \left( \hat S_0^y-\hat S_1^y+\hat S_2^y-\hat S_3^y
-\hat S_0^z+\hat S_1^z+\hat S_2^z-\hat S_3^z\right) ^2\; \; \; &&\nonumber \\
+\left( \hat S_0^z-\hat S_1^z-\hat S_2^z+\hat S_3^z
-\hat S_0^x-\hat S_1^x+\hat S_2^x+\hat S_3^x\right) ^2\; \; \; &&\nonumber \\
+\left( \hat S_0^x+\hat S_1^x-\hat S_2^x-\hat S_3^x
-\hat S_0^y+\hat S_1^y-\hat S_2^y+\hat S_3^y\right) ^2\Big] &&
\end{eqnarray}
This provides a natural model which is minimised by
\begin{eqnarray}
\hat S_0^x+\hat S_1^x-\hat S_2^x-\hat S_3^x=\hat S_0^y-\hat S_1^y+\hat S_2^y
-\hat S_3^y&&\nonumber \\=\hat S_0^z-\hat S_1^z-\hat S_2^z+\hat S_3^z=A\; \;
\; \; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
There are many solutions to this, including the
natural dipolar spirals (as found in Gd$_2$Sn$_2$O$_7$\cite{11}), the cubic
symmetric triple-{\bf q} state and appropriate superpositions.
We now move on to exchange interactions. Susceptibility experiments suggest
that there is a fairly strong antiferromagnetic interaction\cite{23}, and for
a single tetrahedron this is
\begin{eqnarray}
H_1=\frac{1}{2}\Big[ \left( \hat S_0^x+\hat S_1^x+\hat S_2^x+\hat S_3^x\right)
^2\; \; \; &&\nonumber \\+\left( \hat S_0^y+\hat S_1^y+\hat S_2^y+\hat S_3^y
\right) ^2\; \; \; &&\nonumber \\+\left( \hat S_0^z+\hat S_1^z+\hat S_2^z+\hat
S_3^z\right) ^2\Big] &&\nonumber \\\equiv \frac{1}{2}{\bf \hat T}.{\bf \hat T}
\; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
defining a total-spin, ${\bf \hat T}$. Although we have chosen to use a pure
isotropic interaction, the true interaction probably depends on the details of
the occupied orbitals and is consequently likely to be more complicated.
Employing our maximal spin projection this reduces to
\begin{equation}
H_1\mapsto \frac{16}{3}J^2\left( H_0-H_3\right)
\end{equation}
and, as expected, conflicts with `two-in and two-out'. In a similar way to
the previous interaction we can consider
\begin{eqnarray}
H_0-H_3=\frac{1}{24}\Big[ \left( \hat S_0^x+\hat S_0^y+\hat S_0^z-\hat S_1^x
+\hat S_1^y+\hat S_1^z\right) ^2\; \; \; &&\nonumber \\+\left( -\hat S_2^x+\hat
S_2^y-\hat S_2^z+\hat S_3^x+\hat S_3^y-\hat S_3^z\right) ^2\; \; \; &&\nonumber
\\+\left( \hat S_0^x+\hat S_0^y+\hat S_0^z+\hat S_2^x-\hat S_2^y+\hat S_2^z
\right) ^2\; \; \; &&\nonumber \\+\left( \hat S_1^x-\hat S_1^y-\hat S_1^z+\hat
S_3^x+\hat S_3^y-\hat S_3^z\right) ^2\; \; \; &&\nonumber \\+\left( \hat S_0^x+
\hat S_0^y+\hat S_0^z+\hat S_3^x+\hat S_3^y-\hat S_3^z\right) ^2\; \; \;
&&\nonumber \\+\left( \hat S_1^x-\hat S_1^y-\hat S_1^z+\hat S_2^x-\hat S_2^y+
\hat S_2^z\right) ^2\Big] &&\nonumber \\
=\frac{1}{16}\sum _{\alpha \beta }\left( {\bf \hat z}_\alpha .{\bf \hat
S}_\alpha -{\bf \hat z}_\beta .{\bf \hat S}_\beta \right) ^2\; \; \; \; \;
\; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
which is the large-J limit. This is minimised by
\begin{eqnarray}
\hat S_0^x+\hat S_0^y+\hat S_0^z=\hat S_1^x-\hat S_1^y-\hat S_1^z\; \; \; \; \;
\; \; \; \; \; \; \; \; \; \; \; &&\nonumber \\=-\hat S_2^x+\hat S_2^y-
\hat S_2^z=-\hat S_3^x-\hat S_3^y+\hat S_3^z=B&&
\end{eqnarray}
providing another natural Hamiltonian. Again there are many solutions to this,
including all spins being both parallel and perpendicular to the natural
crystallographic directions.
Applying the minimal spin projection to $H_1$ provides
\begin{eqnarray}
H_1\mapsto \left( J+\frac{1}{2}\right) ^2H_1+\frac{4}{3}\left( J-\frac{1}{2}
\right) ^2\left( H_0-H_3\right) &&\nonumber \\-\left( J^2-\frac{1}{4}\right)
H_2\; \; \; \; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
where
\begin{equation}
H_2={\bf \hat T}.\sum _\alpha {\bf \hat z}_\alpha \left( {\bf \hat z}_\alpha .
{\bf \hat S}_\alpha \right)
\end{equation}
Physically we see that the Heisenberg interaction remains dominant, but
there are also terms equivalent to the maximal spin projector and additionally
a new interaction, $H_2$, that would vanish for an antiferromagnet. Note that
the full dipolar interaction, restricted to a single tetrahedron, is
proportional (up to a constant) to
\begin{equation}
H_4+6H_0-3H_2+\frac{7}{3}H_1
\end{equation}
and so all these interactions are `expected'.
Our dipolar arguments naturally lead to $H_0$, $H_1$, $H_2$, $H_4$ and (through
pseudo-spin projection) $H_3$, but our use of the Heisenberg interaction is
less natural. There are likely to be uncontrolled orbitally driven distortions
to the exchange process in this material. We are consequently forced into
treating our Hamiltonian as phenomenological. Note that if for each exchange
bond we simply enhance the bond when the spins are in the Cartesian plane of
the bond, at the expense of when the spins are perpendicular to the plane, then
we generate
\begin{eqnarray}
H=(1-\delta )\left[ \hat S_0^x\hat S_1^x+\hat S_2^x\hat S_3^x\right] \; \; \;
\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \;
&&\nonumber \\+(1+\delta )\left[ \hat S_0^y\hat S_1^y+\hat S_2^y\hat S_3^y+\hat
S_0^z\hat S_1^z+\hat S_2^z\hat S_3^z\right] \; &&\nonumber \\+(1-\delta )\left[
\hat S_0^y\hat S_2^y+\hat S_1^y\hat S_3^y\right] \; \; \; \; \; \; \; \;
\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; &&\nonumber \\+(1+\delta
)\left[ \hat S_0^z\hat S_2^z+\hat S_1^z\hat S_3^z+\hat S_0^x\hat S_2^x
+\hat S_1^x\hat S_3^x\right] &&\nonumber \\+(1-\delta )\left[
\hat S_0^z\hat S_3^z+\hat S_1^z\hat S_2^z\right] \; \; \; \; \; \; \; \; \;
\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; &&\nonumber \\+(1+\delta
)\left[ \hat S_0^x\hat S_3^x+\hat S_1^x\hat S_2^x+\hat S_0^y\hat S_3^y
+\hat S_1^y\hat S_2^y\right] &&\nonumber \\=H_1-\delta H_4-2(1-\delta )S^2
\; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
as desired in the next section. The spin-orbit interaction can easily bias the
exchange, which is mediated by orbital overlaps, in this way.
One experimental observation that is problematic for the state currently
proposed in the literature is the severely reduced moment. If the spins
ordered in the proposed directions, perpendicular to the natural
crystallographic axes, then one would expect to observe essentially the full
moment, whereas a modest ordered moment fits the neutron Bragg data. The
previously proposed explanation for this small moment is quantum fluctuations,
but our prediction has different physics. We suggest that the spins align
quite close to the natural crystallographic directions, where the pseudo-spin
states have a modest observable moment, as suggested by Eq.(\ref{perp}), in
agreement with the experiments.
\section{Spinwaves}
We elect to use the Hamiltonian
\begin{equation}
H=J\sum _t\left[ H_1-\delta H_4-\eta H_0+\xi (H_0-H_3)\right]
\end{equation}
where $t$ labels the tetrahedra, $J$ is the natural energy scale, $\delta $,
$\eta $ and $\xi $ are scale-free parameters. We choose to use {\it positive}
values for these parameters (to stabilise the multiple-{\bf q} states) and
consequently one should probably view this Hamiltonian as phenomenological.
The residual crystal-field interaction can account for $\eta $ and $\xi $, but
the choice of $\delta $ is physically opposite from the presumed dipolar
source for this interaction.
This unexpected `failure' of the dipolar interaction has direct experimental
consequences. In real-space, point-size moments on a lattice have a dipolar
energy proportional to
\begin{equation}
{\cal E}_D=-\frac{1}{2}\sum _{ii'}{\bf \hat S}_i.\frac{\partial }{\partial {\bf
x}}{\bf \hat S}_{i'}.\frac{\partial }{\partial {\bf x}}V\left( \mid {\bf x}
\mid \right) \vert _{{\bf x}={\bf R}_i-{\bf R}_{i'}}
\end{equation}
where $V(X)$ is the Coulomb potential between all lattice sites but vanishes
as $X\mapsto 0$. In reciprocal space this transforms to
\begin{equation}
{\cal E}_D=\frac{1}{2}\sum _{\bf k}\sum _{\bf G}\mid \left( {\bf k}+{\bf G}
\right) .{\bf \hat S}_{\bf k}\mid ^2\tilde V(\mid {\bf k}+{\bf G}\mid )
\end{equation}
where $\tilde V(K)$ is the Fourier transform of $V(X)$. Now since $V(X)$ is
best represented as the Coulomb interaction multiplied by an increasing
function, $\tilde V(K)$ corresponds to the Coulomb interaction multiplied by
a reducing function. To minimise the dipolar interaction we should minimise
the contributions from the Bragg spots, ${\bf G}$, closest to the origin. This
means that the spin-density for these spots should be perpendicular to their
reciprocal-space position. Interestingly, because neutrons scatter with
essentially an equivalent dipolar interaction, this means that these spots
should be {\it maximal} losing nothing from the orientational factor in
neutron scattering\cite{24}. In Gd$_2$Sn$_2$O$_7$, and to a slightly lesser
extent in Gd$_2$Ti$_2$O$_7$, these Bragg spots are indeed maximal\cite{11},
whereas in Er$_2$Ti$_2$O$_7$ they are either small or almost
vanishing. This tells us that the dipolar energy is intrinsically frustrated
and explains why we are forced into using the apparently unphysical sign for
$H_4$.
Since the system exhibits long-range order, we elect to solve our Hamiltonian
in the semi-classical limit to assess the expected style of order and
associated magnons. We employ the Holstein-Primakoff transformation\cite{25}
\begin{equation}
\left[ \begin{matrix} \hat S^x\cr \hat S^y\cr \hat S^z\cr \end{matrix}\right]
\mapsto \left[ \begin{matrix} \left[ \frac{S}{2}\right] ^\frac{1}{2}\left(
b^\dagger +b\right) \cr \left[ \frac{S}{2}\right] ^\frac{1}{2}\left( b^\dagger
-b\right) i\cr S-b^\dagger b\cr \end{matrix}\right]
\end{equation}
with the $z$-axis oriented along the appropriate quantisation direction for
each spin. We start out solving the case $\eta $=0=$\xi $ to find that the
classical solution of Fig.\ref{fig:1} denotes the ground-state manifold. This
then provides
\begin{equation}
{\bf \hat S}_0=\left[ \begin{matrix} {\bf \hat x}&{\bf \hat y}&{\bf \hat z}\cr
\end{matrix}\right] \left[ \begin{matrix}\cos \theta \cos \phi &-\sin \phi
&\sin \theta \cos \phi \cr \cos \theta \sin \phi &\cos \phi &\sin \theta \sin
\phi \cr -\sin \theta &0&\cos \theta \cr \end{matrix} \right] \left[
\begin{matrix} \hat S_0^x\cr \hat S_0^y\cr \hat S_0^z\cr \end{matrix}\right]
\end{equation}
\begin{equation}
{\bf \hat S}_1=\left[ \begin{matrix} {\bf \hat x}&{\bf \hat y}&{\bf \hat z}\cr
\end{matrix}\right] \left[ \begin{matrix}\cos \theta \cos \phi &-\sin \phi
&\sin \theta \cos \phi \cr -\cos \theta \sin \phi &-\cos \phi &-\sin \theta
\sin \phi \cr \sin \theta &0&-\cos \theta \cr \end{matrix} \right] \left[
\begin{matrix} \hat S_1^x\cr \hat S_1^y\cr \hat S_1^z\cr \end{matrix}\right]
\end{equation}
\begin{equation}
{\bf \hat S}_2=\left[ \begin{matrix} {\bf \hat x}&{\bf \hat y}&{\bf \hat z}\cr
\end{matrix}\right] \left[ \begin{matrix}-\cos \theta \cos \phi &\sin \phi
&-\sin \theta \cos \phi \cr \cos \theta \sin \phi &\cos \phi &\sin \theta \sin
\phi \cr \sin \theta &0&-\cos \theta \cr \end{matrix} \right] \left[
\begin{matrix} \hat S_2^x\cr \hat S_2^y\cr \hat S_2^z\cr \end{matrix}\right]
\end{equation}
\begin{equation}
{\bf \hat S}_3=\left[ \begin{matrix} {\bf \hat x}&{\bf \hat y}&{\bf \hat z}\cr
\end{matrix}\right] \left[ \begin{matrix}-\cos \theta \cos \phi &\sin \phi
&-\sin \theta \cos \phi \cr -\cos \theta \sin \phi &-\cos \phi &-\sin \theta
\sin \phi \cr -\sin \theta &0&\cos \theta \cr \end{matrix} \right] \left[
\begin{matrix} \hat S_3^x\cr \hat S_3^y\cr \hat S_3^z\cr \end{matrix}\right]
\end{equation}
for any particular tetrahedron. Bloch transforming provides the excitations
as
\begin{equation}
H=2JS\sum _{{\bf k}\in K_+}\left[ \begin{matrix}{\bf b}^\dagger _{\bf k}&
{\bf b}_{-{\bf k}}\cr \end{matrix}\right] \left[ \begin{matrix}A_{\bf k}&
B_{\bf k}\cr B^\dagger _{\bf k}&A_{\bf k}\cr \end{matrix}\right] \left[
\begin{matrix}{\bf b}_{\bf k}\cr {\bf b}^\dagger _{-{\bf k}}\cr \end{matrix}
\right]
\end{equation}
where we are using the sublattice degrees of freedom as an implicit vector
index and where $K_+$ denotes {\it half} of reciprocal-space, having divided
out inversion. The matrices are explicitly
\begin{widetext}
\begin{equation}
A_{\bf k}=\left[ \begin{matrix} 1+3\delta &-(\sin ^2\theta \cos ^2\phi
+\delta )C_{y+z}&-(\sin ^2\theta \sin ^2\phi +\delta )C_{z+x}&-(\cos ^2\theta
+\delta )C_{x+y}\cr -(\sin ^2\theta \cos ^2\phi +\delta )C_{y+z}&1+3\delta &
-(\cos ^2\theta +\delta )C_{x-y}&-(\sin ^2\theta \sin ^2\phi +\delta )C_{z-x}
\cr -(\sin ^2\theta \sin ^2\phi +\delta )C_{z+x}&-(\cos ^2\theta +\delta )
C_{x-y}&1+3\delta &-(\sin ^2\theta \cos ^2\theta +\delta )C_{y-z}\cr
-(\cos ^2\theta +\delta )C_{x+y}&-(\sin ^2\theta \sin ^2\phi +\delta )C_{z-x}&
-(\sin ^2\theta \cos ^2\theta +\delta )C_{y-z}&1+3\delta \cr
\end{matrix}\right]
\end{equation}
\begin{equation}
B_{\bf k}=\left[ \begin{matrix} 0&(\cos \theta \cos \phi -i\sin \phi )^2C_{y+z}
&(\cos \theta \sin \phi +i\cos \phi )^2C_{z+x}&\sin ^2\theta C_{x+y}\cr
(\cos \theta \cos \phi -i\sin \phi )^2C_{y+z}&0&\sin ^2\theta C_{x-y}&
(\cos \theta \sin \phi +i\cos \phi )^2C_{z-x}\cr (\cos \theta \sin \phi +i\cos
\phi )^2C_{z+x}&\sin ^2\theta C_{x-y}&0&(\cos \theta \cos \phi -i\sin \phi )^2
C_{y-z}\cr \sin ^2\theta C_{x+y}&(\cos \theta \sin \phi +i\cos \phi )^2C_{z-x}&
(\cos \theta \cos \phi -i\sin \phi )^2C_{y-z}&0\cr \end{matrix}\right]
\end{equation}
\end{widetext}
where
\begin{equation}
C_{\alpha \pm \beta }=\cos (k_\alpha \pm k_\beta )
\end{equation}
We can now diagonalise this Hamiltonian with a Bogoliubov transformation to
construct the spinwave spectrum, $E_{{\bf k}\alpha }$. In Fig.\ref{fig:2}
\begin{figure}
\includegraphics[height=7.2 cm, width=8.4 cm]{ertif2}
\caption{\label{fig:2} The four spinwave branches for the case $\delta $=0.0001
in units of $2JS$, parallel to one of the Cartesian axes for the case
corresponding to the experimentally proposed ground-state.}
\end{figure}
we look at the case where $\delta $ is very small. There are two almost flat
bands on the energy scale of $2JS\surd \delta $. These control the original
degeneracy of the Heisenberg model which has been lifted by the inclusion of
$H_4$. In Fig.\ref{fig:3}
\begin{figure}
\includegraphics[height=7.2 cm, width=8.4 cm]{ertif3}
\caption{\label{fig:3} The four spinwave branches for the case $\delta $=0.2
in units of $2JS$, parallel to one of the Cartesian axes for the case
corresponding to the experimentally proposed ground-state.}
\end{figure}
we take the case of a small but relevant $\delta $.
We see that the flat bands have now developed a little dispersion and
are close to the highest energy excitations. This plot has much in common
with the experiments\cite{2}, but also has some clear problems. The low energy
mode is clearly visible, as is the almost flat high energy mode, but instead of
a linear dispersion at low energy we find a quadratic dispersion. This is
because our model is actually degenerate across all the multiple-{\bf q}
states of Fig.\ref{fig:1}, but if the particular ground-state were picked
out by the energetics then the spinwave would be expected to harden.
We are claiming that the low energy mode observed in the experiment is a rare
form of longitudinal spinwave and so we now take time to explain this idea.
We start out with an isotopic single-{\bf q} state, $\theta $=0, oriented along
the $z$-axis. The magnetic scattering would then appear at (220), but the
(022) and (202) spots would be absent. The usual Goldstone modes would be
expected close to (002) and analogous spots on the body-centre-cubic
superlattice generated by $\{ (2\bar 2\bar 2), (\bar 22\bar 2), (\bar 2\bar 22)
\} $, but there would usually be a spinwave gap at (200) and (020), as these
points are non-magnetic and unrelated to (002) by any magnetic symmetry. The
usual Goldstone modes are transverse; they are associated with the original
state being tilted off axis and correspond to a small amount of magnetism
at the same {\bf q}-point but in a perpendicular direction. The initially
gapped modes at (200) and (020) correspond to small magnetic distortions of
the other two styles of point-group related magnetism which are currently not
present in the system. Physically they amount to small changes in the $\phi $
and $\theta $ of Fig.\ref{fig:1}. The softening of these modes then
corresponds to a {\it phase transition} from the original single-{\bf q} state
into a multiple-{\bf q} state. The incoming state naturally uses $\phi $ to
become a double-{\bf q} magnet or $\theta $ to become a triple-{\bf q}
magnet\cite{4}. In a normal rare-earth material the spin-orbit interaction
opens a huge gap in the standard transverse mode, but the longitudinal mode,
which links to the other magnetic states can be found at low energies, as in
CeAs\cite{18}. The physical idea is then that as the angle $\theta $ smoothly
rotates, from zero at the single-{\bf q} state to $\cos \Theta =\frac{1}{\surd
3}$ in the cubic triple-{\bf q} state, the mode remains at zero energy and then
generates a gap again in the cubic state. Including the parameter $\eta $ into
the spinwave calculation then stabilises the cubic triple-{\bf q} state and
provides the gapped spin-wave dispersion of Fig.\ref{fig:4}.
\begin{figure}
\includegraphics[height=7.2 cm, width=8.4 cm]{ertif4}
\caption{\label{fig:4} The four spinwave branches for the case $\delta $=0.2,
and $\eta $=0.2, which stabalises the cubic triple-{\bf q} state ($\phi
=\frac{\pi }{4}$ and $\theta =\Theta $), in units of $2JS$, parallel to one of
the Cartesian axes.}
\end{figure}
We now take time to offer an exactly solvable model which highlights some of
the issues
\begin{equation}
H=-\delta H_4+\xi \left( H_0-H_3\right)
\end{equation}
This model also has the multiple-{\bf q} states of Fig.\ref{fig:1} as its
ground-state manifold, but it has the rather elementary dynamics controlled by
\begin{equation}
H=2JS\left[ \begin{matrix} \left[ \delta +\frac{\xi }{24}Z^*Z\right] C_{\bf k}&
\frac{\xi }{24}ZZC_{\bf k}\cr \frac{\xi }{24}Z^*Z^*C_{\bf k}&\left[ \delta
+\frac{\xi }{24}Z^*Z\right] C_{\bf k}\cr \end{matrix}\right]
\end{equation}
where
\begin{equation}
C_{\bf k}=\left[ \begin{matrix} 3&-C_{y+z}&-C_{z+x}&-C_{x+y}\cr -C_{y+z}&3&
-C_{x-y}&-C_{z-x}\cr -C_{z+x}&-C_{x-y}&3&-C_{y-z}\cr -C_{x+y}&-C_{z-x}&
-C_{y-z}&3\cr \end{matrix} \right]
\end{equation}
and
\begin{equation}
Z=\cos \theta \sin \phi +i\cos \phi +\cos \theta \cos \phi -i\sin \phi -\sin
\theta
\end{equation}
The spinwave dispersion is given by
\begin{widetext}
\begin{equation}
E_{{\bf k}\alpha }=2JS\left[ \delta \left( \delta +\frac{\xi }{12}\left[ \left(
\sin \theta \sin \phi -\cos \theta \right) ^2+\left( \cos \theta -\sin \theta
\cos \phi \right) ^2+\left( \sin \theta \cos \phi -\sin \theta \sin \phi
\right) ^2\right] \right) \right] ^\frac{1}{2}\epsilon _{{\bf k}\alpha }
\end{equation}
\end{widetext}
where
\begin{equation}
\epsilon _{{\bf k}\alpha }=4,4,2\pm \left[ 1+3\gamma _{\bf k}\right]
^\frac{1}{2}
\end{equation}
and $\gamma _{\bf k}$ is the face-centre-cubic structure factor for the
underlying periodicity. The quantity $\epsilon _{{\bf k}\alpha }$ is
essentially the pyrochlore lattice structure factor, exhibiting a gapless
band and two flat bands describing the degeneracy (clear at finite energy).
This model (ignoring the quadratic dependence at low energy) is enough
to describe the observed inelastic neutron scattering\cite{2}, and hence this
experiment does not clearly shed much light on {\it which} multiple-{\bf q}
state might actually be stable.
Although all our multiple-{\bf q} states are classically degenerate, the
quantum fluctuations lift this degeneracy\cite{26} and tend to stabilise
collinear states. For the current exactly solvable model we can calculate
this quantum fluctuation energy analytically. At zero temperature we find that
the fluctuation energy per spin, $Q$, is
\begin{equation}
Q=3JS\left( -\delta -\frac{\xi }{24}\mid Z\mid ^2+\left[ \delta \left( \delta
+\frac{\xi }{12}\mid Z\mid ^2\right) \right] ^\frac{1}{2}\right)
\end{equation}
where
\begin{eqnarray}
\mid Z\mid ^2=3-\left( \sin \theta \cos \phi +\sin \theta \sin \phi +\cos
\theta \right) ^2\; \; \; \; \; \; \; &&\nonumber \\=\left( \sin \theta \sin
\phi -\cos \theta \right) ^2+\left( \cos \theta -\sin \theta \cos \phi \right)
^2&&\nonumber \\+\left( \sin \theta \cos \phi -\sin \theta \sin \phi \right) ^2
\; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; \; &&
\end{eqnarray}
This fluctuation energy vanishes for the cubic triple-{\bf q} state and is minimised when the
spins are perpendicular to the natural crystallographic directions. This
exactly solvable model has very anisotropic interactions which do not promote
standard collinear states. If we accept that our spin is pseudo-spin half
and that the quantum fluctuation energy is of the same order as the classical
energy, then we can now attempt a physical explanation for the system. If
there is a classical energy promoting the cubic triple-{\bf q} state, such as
$H_0$ or $H_3$, then this could balance the quantum fluctuation energy at the
observed experimental state. However, careful analysis of the current model
shows that the state jumps discontinuously from the cubic state to one with the
spins perpendicular to the crystallographic directions and so the current model
is too simplistic.
Returning to our more physical model, $H_1-\delta H_4$, we can calculate the
quantum fluctuation energy as
\begin{equation}
Q=JS\left( -1-3\delta +\frac{1}{4N}\sum _{{\bf k}\alpha }E_{{\bf k}\alpha }
\right)
\end{equation}
at zero temperature, and we offer an example in Fig.\ref{fig:5}.
\begin{figure}
\includegraphics[height=7.2 cm, width=8.4 cm]{ertif5}
\caption{\label{fig:5} The quantum fluctuation energy for the case $\delta
$=0.2, as a function of $\theta $ (in units of $\frac{\pi }{2}$)with
$\phi $=$\frac{\pi }{4}$ (solid curve). We have included a small $\eta $ to
achieve the blue curve.}
\end{figure}
As expected, the ground-state is now the single-{\bf q} state (black curve). We next add
a carefully scaled classical contribution favouring the cubic state and now an
intermediate state appears relatively stable (blue curve). However the energetics are not very
`stable' and rather than continuing to evolve smoothly, the ground-state tends to make a small
discontinuous jump when close to the cubic state, to the cubic state.
Having analysed a particular case, we now move on to generalities. Our
complete Hamiltonian is
\begin{eqnarray}
H=J\sum _t\Big[ \alpha H_1-\delta H_4-\eta H_0+\xi (H_0-H_3)&&\nonumber \\
-\frac{3}{2}\beta H_2+\epsilon H_5\Big] \; \; \; \; \; \; \; \; &&
\end{eqnarray}
where $H_5$ is a local interaction which lifts the azimuthal symmetry and
provides a crystal-field preference appropriate to the weak hexagonal symmetry
around the natural crystallographic directions. We can semi-classically
solve such an interaction in general. There are four natural ground-states:
the dipolar spirals\cite{11}, a generalised `two-in and two-out' state (typical
for spin-ice\cite{7}), the pure triple-{\bf q} state\cite{4} and finally a
multiple-{\bf q} state with the spins perpendicular to the natural
crystallographic directions, but pointing along one of the twelve natural
directions allowed by the weak hexagonal interaction. The key observation is
that in {\it all} of these phases there is a natural gap in the spinwave
spectrum. Our generic Hamiltonian is very anisotropic and that generates
the expected loss of continuous symmetry and a consequent gap. The observed
gapless state is truly unexpected.
There is no easy way to eliminate the gap in the spiral and `two-in and
two-out' phases, but since Er$_2$Ti$_2$O$_7$ has a multiple-{\bf q}
ground-state this is irrelevant. One way to close the gap is to omit the
hexagonal crystal-field interaction, $\epsilon \mapsto 0$. This is the current
core of the picture prevalent in the literature\cite{1,2,3,15}. If it were
exact, then one could use a magnetic field to access this degeneracy but the
current experiments point to the particular azimuthal angle being pinned and
nonreactive to a field. A weak hexagonal field provides a small gap. However it is
possible that the observed linear dispersion, and specific heat, have not been accessed at low
enough energies, and temperatures respectively, to observe this gap.
Another way to close the gap, our explanation, is that the
multiple-{\bf q} degeneracy is active. We would leave the hexagonal
interaction, which stabilises $\phi =\frac{\pi }{4}$, but omit the excess
local crystal-field term, $\eta \mapsto 0$. The inclusion of the hexagonal
interaction removes one inconsistency in our model; the quadratic dispersion at
low energy becomes linear as required, and our continuous degeneracy is in the
angle $\theta $ and not $\phi $. We believe that once the particular choice
of $\theta $ is chosen by some collective compromise, the gaplessness still
remains. Since there is no such collective competition in our semi-classical
analysis, we have resorted to suggesting that quantum fluctuations could
provide such an asymmetric $\theta $, without offering any rigorous
calculations to prove this.
\section{Conclusions}
Rare-earth magnets are usually very anisotropic with a gap to magnetic
excitations. Er$_2$Ti$_2$O$_7$ is an unusual example of a gapless rare-earth
magnet. The experimental evidence points towards a multiple-{\bf q}
antiferromagnet where the spins do not point along crystallographically natural
directions. The low energy mode then smoothly varies this incommensurate
angle. The single-domain single-crystal measurements indicate that the angles
$\phi $=$\frac{\pi }{4}$ and (the incommensurate) $\theta \sim
\frac{\pi }{5}$ are selected.
The observed spinwave spectrum is well represented by the natural pyrochlore
structure factor and hence does not restrict the model much, except that the
spectrum is gapless and so a continuous degeneracy is required. The observed
orientational factors in the neutron scattering are contradictory to the
expectations of the dipolar interaction and this leads us to propose the model
with the opposite sign for $H_{4}$ in order to fit the experiments.
Our physical insight arises from the balance between quantum fluctuations and
classical interactions. Our Hamiltonian, $H_1$-$\delta H_4$, is classically
degenerate, with the states of Fig.\ref{fig:1} as the ground-state manifold.
The quantum fluctuations stabilise the single-{\bf q} state, and because the
pyrochlore lattice is highly frustrated the quantum fluctuation energy is
expected to be large. Due to the strong spin-orbit and crystal-field effects
there are additional interactions which happen to favour the natural
crystallographic directions. If we imagine increasing such an interaction
smoothly from zero, at some strength it will overturn the quantum fluctuation
energy and drive the system into a multiple-{\bf q} state. We are proposing
that this transition would be second-order and would pass through the
intermediate gapless phases where $\theta $ varies from zero to $\Theta $, with
$Er_{2}Ti_{2}O_{7}$ part-way through this process at the observed
value of $\theta \sim \frac{\pi }{5}$.
Our proposal requires a multiple-{\bf q} state with an unnatural angle
$\theta $ which is not clearly derived nor convincingly explained. The current
picture in the literature has the natural angles $\theta $=$\frac{\pi }{2}$ and
$\phi $=$\frac{\pi }{4}$ and so would be expected to have a gap.
This gap needs to be too small to be observed. This state should also have
a dominant (111) spot, as this Bragg spot structure factor is perpendicular
to its momentum transfer, whereas the spot is observed to be small. The only
weakness in our argument is the form-factor, which would have to have a severe
and unexpected orientational dependence.
There are now two plausible states that will fit the experiments: the original
idea\cite{15}, with spins perpendicular to the natural crystallographic
directions, and our proposal with a non-symmetric angle $\theta $. How can
we decide between them experimentally? The original proposal should exhibit
a gap, controlled by the hexagonal crystal field, which can be sought. If
this gap is tiny, then an external magnetic field should be able to rotate
the angle $\phi $ and dramatically change the magnetic Bragg scattering, with
the ratio of intensities changing from $(1,1,4)$ to $(3,3,0)$ for the three
main Bragg spots. One would also need an explanation for why the $(111)$-type
Bragg spots were so small, either from dramatic form-factor dependence or from
some other as yet unresolved experimental issue. Our proposal has rather
different issues. Since we fit the known data the experiments are not too
pertinent, it is the theory that is critical. We need a model that stabilises
the unsymmetric state and we need the state to remain gapless, both issues
having been side-stepped in this article.
Our proposed model is a pseudo-spin half model on the pyrochlore lattice but
our semi-classical magnetic techniques are not sophisticated enough to predict
our expectations. Better theoretical analysis of the quantum mechanical model
is required to make a more detailed comparison with experiment.
\begin{acknowledgments}
We wish to acknowledge useful discussions with E.A. Blackburn.
\end{acknowledgments}
|
\section{Introduction}
The measurement of magnetic fields harbor several difficulties due to
the fact that they are a vector quantity. In order to measure the
absolute magnitude and direction of the magnetic field, it is
necessary to sample the two components in the plane of the sky and the
line-of-sight component. While dust polarization can measure the
direction of a magnetic field in the plane of the sky, the
line-of-sight component intensity can be directly measured. This is
accomplished through the detection of the normal Zeeman effect which
occurs when an atom or molecule is in the presence of a magnetic
field. Most transitions have a Zeeman splitting factor that is too
small to measure unless in dense, spatially-compact environments with
exceedingly large magnetic field strengths. However, HI, OH, and CN
have splitting factors that are large enough to measure the weaker
magnetic fields that are expected around sites of active star
formation in molecular clouds. The CN N = 1-0 transition has seven
strong hyperfine components (Table \ref{cntable}) that have a large
distribution of Zeeman splitting factors. The two strongest hyperfine
components, with relative strengths of 27 and 10 respectively, also
have large splitting factors.
\begin{deluxetable}{lcccc}
\tablecaption{CN Hyperfine Components\label{cntable}}
\tablehead{\colhead{Molecule} & \colhead{Transition} & \colhead{$\nu$ (MHz)}
& \colhead{Z (Hz/$\mu$G)\tablenotemark{\emph{a}}}
& \colhead{RI\tablenotemark{\emph{b}}}}
\tablewidth{0pt}
\startdata
CN & N=$1-0$ J=$\frac{1}{2}-\frac{1}{2}$ F=$\frac{1}{2}-\frac{1}{2}$
& 113123.337 & --\tablenotemark{\emph{c}} & 1 \\
CN & N=$1-0$ J=$\frac{1}{2}-\frac{1}{2}$ F=$\frac{1}{2}-\frac{3}{2}$
& 113144.192 & 2.18 & 8 \\
CN & N=$1-0$ J=$\frac{1}{2}-\frac{1}{2}$ F=$\frac{3}{2}-\frac{1}{2}$
& 113170.528 & -0.31 & 8 \\
CN & N=$1-0$ J=$\frac{1}{2}-\frac{1}{2}$ F=$\frac{3}{2}-\frac{3}{2}$
& 113191.317 & 0.62 & 10 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{3}{2}-\frac{1}{2}$
& 113488.140 & 2.18 & 10 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{5}{2}-\frac{3}{2}$
& 113490.982 & 0.56 & 27 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{1}{2}-\frac{1}{2}$
& 113499.639 & 0.62 & 8 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{3}{2}-\frac{3}{2}$
& 113508.944 & 1.62 & 8 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{1}{2}-\frac{3}{2}$
& 113520.414 & --\tablenotemark{\emph{c}} & 1 \\
\enddata
\tablenotetext{\emph{a}}{Z is the Zeeman Splitting Factor for each hyperfine
component of CN.}
\tablenotetext{\emph{b}}{RI is the relative intensity of each component.}
\tablenotetext{\emph{c}}{The Z's for these hyperfine components were not calculated due to their low RI.}
\tablerefs{Falgarone et al. 2008}
\end{deluxetable}
\citet{falgarone08} surveyed 14 molecular cloud cores and measured the
line-of-sight magnetic field strength using the CN Zeeman effect. The
strongest magnetic field detected was $1.10\pm 0.33$ mG towards the
Ultra-Compact HII (UCHII) region of W3(OH) \citep{hoare04}, a high
mass star formation site in the W3 molecular cloud complex. For these
measurements, \citet{falgarone08} used the IRAM 30-meter
telescope. Due to the relatively large beam size of the IRAM-30 meter
telescope (23\arcsec\ or 46,000 AU in diameter at the 2 kpc distance
\citep{hachisuka06} of W3(OH)), only very limited mapping to find the
emission peak was performed. From these data, it is not possible to
determine the structure of the magnetic field or which material the
field is associated with.
W3(OH) contains a high-mass O star and has been extensively studied
over the years since the discovery of several OH maser sites within it
\citep{raimond69}. A star forming region ~6\arcsec\ east of W3(OH) was
later discovered by \citet{turner84} in HCN. This smaller region was
also studied for maser activity and found to contain several H$_2$O
masers \citep{wynnwilliams72}. While this second source was initially
not observed in the continuum, with the development of more sensitive
instruments, it was eventually detected in dust continuum studies
\citep{wilner95,wyrowski97}. Within the last few years, further high
resolution studies with instruments such as BIMA have led to the
detection of multiple dense cores believed to be sites of active star
formation within the Turner-Welch Object \citep{chen06}.
In order to determine the spatial distribution of the magnetic field,
an understanding of the CN gas is needed. There is some question,
however, as to the exact physical conditions that CN
traces. \citet{hilyblant08} used the IRAM 30-m telescope to conclude
that CN in prestellar cores stays in the gas phase at densities close
to $10^6$ cm$^{-3}$ and can serve as a kinematic tracer of high
density gas. They compared their CN maps with N$_2$H$^+$, another high
density tracer, and found the results to be in good
agreement. Therefore, it is expected that the measured magnetic field
strength derived from CN measurements would be in gas that is
associated with high density regions of active star
formation. However, before detailed conclusions can be drawn from
these magnetic field measurements, a high angular resolution study of
the CN gas distribution is required. An ideal instrument for this is
an interferometer such as CARMA, as its high spatial resolution and
ability to sample baselines as short as 6m allow us to sample material
both spatially compact and widespread.
\section{CARMA Observations}
With CARMA \citep{bock06} it is possible to achieve a resolution of
approximately 2.5$\arcsec$ by combining maps generated with the C, D,
and E arrays. By using three different array configurations, we have
the ability to probe small scale structure while retaining the ability
to image large scale features that would otherwise be resolved out in
the C and D arrays. CARMA currently does not have polarization
capability and therefore cannot perform magnetic field measurements;
however, by mapping these regions at high resolution we can gain
crucial information as to the structure of the regions that are being
sampled by the IRAM observations. To assist with this, we use the
advanced features of CARMA's correlator to sample several spectral
lines simultaneously, so we can compare the structure of the CN
emission with emission of other well studied tracer molecules. We
will also gain information on which regions are possible to map in the
Zeeman effect at high resolution and for long integration times with
an instrument capable of making dual-polarization measurements, when
such an interferometer array becomes available.
The CARMA observations began in Spring 2007 and were completed in Fall
of 2009. About 66 hours of observing time was used to map a 1
arcmin$^2$ region around W3(OH). This time was used to observe the
following molecular tracers: CN, C$^{18}$O, N$_2$H$^+$, HCN, and
HCO$^+$ (Table \ref{transitionstable}). Due to the design of the CARMA
correlator and the relatively large frequency range within the 3-mm
band that these tracers span, the measurements had to be performed in
two separate tracks: one for CN and C$^{18}$O, and a second for
N$_2$H$^+$, HCN, and HCO$^+$. Approximately 50 hours (Table
\ref{w3ohtime}) was spent for the CN band and 16.2 hours was spent for
the N$_2$H$^+$ band (not counting additional time spent in B array,
discussed below). In order for us to reach our signal to noise goal,
the tracks containing N$_2$H$^+$ required much less observing
time. Table \ref{w3ohtime} shows the observing array details. Our
pointing center is located on the continuum source, an ultra-compact
HII region, located at: 02:27:03.7 RA, +61:52:25 DEC (J2000). This is
slightly offset from the IRAM pointing position.
\begin{deluxetable}{lcr}
\tablecaption{Observed Transitions\label{transitionstable}}
\tablehead{\colhead{Molecule} & \colhead{Transition} & \colhead{$\nu$ (MHz)}}
\tablewidth{0pt}
\startdata
HCN & N=$1-0$ F=$1-1$ & 88630.415 \\
HCN & N=$1-0$ F=$2-1$ & 88631.847 \\
HCN & N=$1-0$ F=$0-1$ & 88633.936 \\
HCO$^+$ & N=$1-0$ & 89188.526 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$1-1$ F=$0-1$ & 93171.621 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$1-1$ F=$2-2$ & 93171.917 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$1-1$ F=$1-0$ & 93172.053 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$2-1$ F=$2-1$ & 93173.480 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$2-1$ F=$3-2$ & 93173.777 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$2-1$ F=$1-1$ & 93173.967 \\
N$_2$H$^+$ & N=$1-0$ F$_1$=$0-1$ F=$1-2$ & 93176.265 \\
C$^{18}$O & N=$1-0$ & 109782.176 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{3}{2}-\frac{1}{2}$ & 113488.140 \\
CN & N=$1-0$ J=$\frac{3}{2}-\frac{1}{2}$ F=$\frac{5}{2}-\frac{3}{2}$ & 113490.982 \\
\enddata
\end{deluxetable}
\begin{deluxetable}{clll}
\tablecaption{W3OH Observation Lengths\label{w3ohtime}}
\tablehead{\colhead{Array Config} & \colhead{CN, C$^{18}$O} & \colhead{HCN, HCO$^+$, N$_2$H$^+$} & \colhead{Resolution}}
\tablewidth{0pt}
\startdata
C & 28.0h & 7.4h & 1.5\arcsec \\
D & 8.1h & 4.3h & 5.0\arcsec \\
E & 11.0h & 4.5h & 10.0\arcsec \\
B & 30h & 12h & $\sim 0.7$\arcsec \\
\enddata
\end{deluxetable}
We produced maps for 3-mm lines of CN, C$^{18}$O, N$_2$H$^+$, HCN, and
HCO$^+$ at a resolution of approximately 2.5$\arcsec$. The composite
CARMA primary beam at half power at this frequency is slightly more
than 60$\arcsec$ in diameter, reflected in our maps, which are
64$\arcsec$ on a side. Since baselines as short as 6m are included,
the maps are sensitive to structure smaller than about 90$\arcsec$, so
that the spatial dynamic range is almost 40:1. At the distance of
W3(OH) (2 kpc), this means the maps are about 0.62 pc across, and the
resolution is about 0.024 pc or 5,000 AU. For all of the spectral line
observations, we used the 8 MHz spectral line mode, which for CN
(N$_2$H$^+$) provides a velocity coverage of $\sim$21 km s$^{-1}$
($\sim$25.8 km s$^{-1}$) and a resolution of 0.335 km s$^{-1}$ (0.415
km s$^{-1}$). The other transitions have similar coverages and
resolutions which vary depending on their rest frequency. The large
velocity coverage is necessary to simultaneously image several
hyperfine components while still maintaining a high spectral
resolution. Our data also contains two 500 MHz continuum bands in each
track giving us four separate continuum windows between 88 GHz and 113
GHz. These continuum maps, while having a lower spatial resolution
than some previously published maps of W3(OH) \citep{chen06}, have
a significantly higher signal-to-noise ratio.
MIRIAD was used for data reduction, specifically with the
modifications for use with CARMA. Due to the nature of CARMA, with 3
distinct primary beams, the data were reduced as a mosaic data set
with a common pointing center. Passband calibration was performed on
all tracks by utilizing the CARMA system noise source. Even though
suitable passband calibrators were observed for every track, it was
decided that better passband solutions could be obtained from the
noise source. Flux calibration was performed on each individual track
with a 15 minute observation of MWC349, Mars, or Uranus, whichever was
best suitable at observation time as the intrinsic flux of these
sources is very well modeled. Since our data were obtained over such a
long period of time, the flux of our primary phase calibrator varied
significantly. We determine the flux of the phase calibrator through a
bootstrapping process by which our flux calibrator measurements are
used to calibrate the flux scale for the point source phase
calibrator. It is estimated that there is an inherent 20\% error with
this flux calibration technique. Phase calibration was performed
primarily on 0359+509 with a secondary calibrator 0102+584, used if
the primary was not visible for a significant portion of the
track. The primary phase calibrator had an averaged bootstrapped flux
of 6.16 Jy over 10 tracks obtained over a period of 20 months. The
flux steadily increased from 3.9 Jy to 10.7 Jy over this time. In
order to corroborate the flux measurements we compared our data with
that obtained by the CARMA flux calibration commissioning task. This
independent flux measurement of 0359+509 is 9.83 Jy (from Fall 2009),
consistent with our measurement of 10.7 Jy from the same time period,
assuming the 20\% uncertainty mentioned above. Gain and phase
calibration was performed with the \emph{gfiddle} routine which fits
an n-th order polynomial to the phase measurements on the phase
calibrator. This method was chosen in contrast to the more widely used
\emph{selfcal} technique (on the calibrator) due to the weak signal
strength of some of our spectral lines. However the benefits of this
technique over the selfcal technique have not been shown; overall this
technique may not have any net advantage. It may be possible to gain
an increased signal to noise by performing self-calibration on W3OH
itself, as the continuum source in the center of the map remains
unresolved in all of the array configurations and is bright enough to
perform phase calibration in the spectral channels. However, the
absolute position of the phase center of the map is not retained and
flux calibration is not guaranteed to succeed. These details are
described in detail below in the description of our B-array data.
Since CARMA is composed of two types of telescopes of differing sizes,
we have to handle the fact that we have multiple primary beam types
associated with our data. CARMA has both 10-meter and 6-meter dishes
which form 3 effective primary beams with the same phase center: one
for the pairing of 10-meter to 10-meter dish, one for 6-meter to
6-meter, and a third of 10-meter to 6-meter. At the CN line frequency
the 6-meter primary beam half power beam width (HPBW) is
$\sim$111\arcsec, the 10-meter HPBW is $\sim$66\arcsec (at the
N$_2$H$^+$ frequency of 93 GHz, these HPBWs are 135\arcsec\ and
81\arcsec, respectively), while the 6m-10m primary beam size is in
between that of the other two. We conservatively constrained our maps
to $\sim$1 arcmin$^2$ in order to easily compare the other maps with
the CN data. The restoring beam size is calculated by fitting the
dirty beam with a Gaussian beam. This is the effective resolution of a
map, and is the value quoted in all of our figures.
Image conversion from the UV-dataset into the spatial domain was
performed by the MIRIAD routine \emph{invert}. System temperature
weighting was used to properly downweight data taken at low elevation
or in the unlikely event that data were improperly flagged. The CARMA
control system has an intricate flagging mechanism that operates if
one of many error conditions are met including: telescope tracking
errors, pointing errors, receiver problems (dewar temperature, LO
frequency, etc), system temperature, and other computer or correlator
errors that could hinder data integrity. All data were additionally
inspected for extreme system temperatures, unphysical antenna gains,
time regions without converging phase solutions, and poor weather
conditions that could influence the quality of the data. In a few
cases (usually in older tracks before some automated flagging
conditions were introduced) data that contained systematic errors were
manually flagged. This additional flagging increased the
signal-to-noise of these few tracks by a factor of 2-3. In the
\emph{invert} step, we weighted our UV data using a Briggs visibility
weighting robustness parameter of 1 \citep{briggs99}, which
provides a slightly increased signal-to-noise ratio over uniform
weighting at the expense of a larger synthesized beam size. The
relatively weak CN emission prompted the weighting of data in this
manner. Since we are detecting low-level emission we did not want
large sidelobes to obscure the detections. If we changed the weighting
more towards Natural weighting, we could have reduced the noise level
further; however, the beam size would be significantly larger and we
would be affected by significant sidelobes. Cleaning was performed
with the MIRIAD task \emph{mossdi}, the mosaic version of the standard
Steer clean routine. We used a multi-step cleaning technique in order
to prevent under-cleaning the source. A small number of clean
iterations was performed in order to remove large scale beam patterns
while leaving much of the low intensity source features
uncleaned. From this initial clean, we ran the MIRIAD task
\emph{restor} which used the clean data to produce a map without beam
effects. We used the average off-line noise level in this map to limit
how much flux was cleaned from a second iteration of \emph{mossdi} and
to prevent overcleaning. This second stage \emph{mossdi} iteration was
followed by another invert step to produce a final, deep cleaned
map. In theory, this process can be repeated ad infinitum, however, if
the noise level calculated in the first step is accurate, and the
maximum number of cleaning iterations is large enough such that we are
guaranteed to clean down to the noise level, this two step process is
adequate.
\section{Results}
\subsection{CN}
Figure \ref{w3ohmap} shows integrated line maps of the five
species, while Figure \ref{cnchannel} shows channel maps. The continuum
source was removed from the spectral line maps by averaging several
off-line channels together and subtracting that from each individual
channel. Towards the continuum source, we see three of the species
strongly in absorption. Besides the Turner-Welch object and the
continuum source, there are two other regions that are notable in
CN. The western side of the map contains very diffuse CN emission,
while the eastern side of the map shows a compact and complex CN
emission source. In order to compare the CARMA CN data with the IRAM
CN data \citep{falgarone08}, we also generated a map where the
CARMA channel data were convolved with a 23$\arcsec$ Gaussian beam,
effectively smoothing the CARMA data to match the IRAM resolution (see
\S 4.2). By comparing these maps, there is evidence that both spectra
are dominated by diffuse emission.
Compared to the other molecules, the CN emission appears to be diffuse
but ``clumpy.'' Under closer scrutiny, it is much more complex. Our 8
MHz spectral window is centered at 113.490982 GHz, the frequency of
CN's strongest hyperfine component. This frequency and window size was
chosen to allow us to simultaneously image a second hyperfine
component of the CN N=1-0 transition at 113.488 GHz in the same
spectral window. These two hyperfine components are separated by 7.9
km s$^{-1}$. Since this window covers approximately 21 km s$^{-1}$ of
velocity space, we expect to fully resolve and image both
lines. Multiple velocity components in the image make it difficult to
determine which hyperfine component some emission belongs to. However,
we have been able to identify at least three distinct velocity
components, representing separate regions around W3(OH) (Figure
\ref{cnchan}).
The north-west and south-west components appear to be very
diffuse. They are centered at approximately -47.25 km s$^{-1}$ and
-44.5 km s$^{-1}$, respectively. It is difficult to completely isolate
these two velocity components due to blending and overlap of the
hyperfine lines. The maps in Figure \ref{cnchan} reflect some of this
``blending.'' It occurs because the second hyperfine component of one
velocity partially overlaps the primary hyperfine component of the
other velocity. For example, the second and third maps in Figure
\ref{cnchan} are cross-contaminated. This is easily visible when the
positions of the emission in the two maps are compared; they have a
significant spatial overlap region.
The north-eastern component is more complex. In addition to containing
diffuse emission, multiple denser clumps are visible. The significance
of some of these clumps are not clear from the integrated line
map. Between the larger southernmost clump and the smaller, northern
clump are a string of small, presumably unresolved clumps only seen in
a single channel each. When viewing an averaged line map, many of
these features are spatially smoothed and are not prominent (as in
Figure \ref{cnchan}). The densest portions of this region are
concentrated in two lobes, not unlike the typical signature of an
outflow. Figure \ref{outflowcm} shows channel maps of the 15 channels
around the strongest hyperfine component. There does not appear to be
any blending as was seen in the south west component, so we are
confident that this channel map is not contaminated. An averaged
contour plot (Figure \ref{outflowcombined}) shows the physical
comparison between the two ``lobes'' of the emission. The top (red
shifted) lobe is averaged from the six channels from -47.6 km s$^{-1}$
to -49.2 km s$^{-1}$, while the bottom (blue shifted) lobe is averaged
from the six channels from -49.2 km s$^{-1}$ to -50.8 km
s$^{-1}$. While this does have the signature of a typical outflow, it
is not accompanied by any coincident emission from any of the other
species, nor is there any accompanying IR emission source visible in
publicly available IR catalogs. Due to the extent of the low level
emission around the whole region, and lack of evidence of an outflow
generating source, it is more likely that this is not an outflow, but
rather bulk rotation of a clump of gas.
\subsection{HCN, HCO$^+$, C$^{18}$O}
The maps for HCN, HCO$^+$, C$^{18}$O all appear to be tracing the same
material, which does not coincide with either the CN or N$_2$H$^+$
emission (see Figures \ref{hcnchannel}-\ref{c18ochannel} for channel
maps). However, the peak position of these spectral lines coincides
with the HCN detection of the Turner-Welch object first reported in
\citet{turner84}. At 2.5\arcsec\ resolution we do not appear to be
resolving the object; previous high resolution (sub-arcsecond) studies
have been conducted in order to resolve this object in the continuum
\citep{chen06}, which resulted in the detection of multiple
cores. Previous HCO$^+$ studies \citep{wink94} had similar spatial
resolution to our data, however, our data has significantly better
signal-to-noise and samples material that was resolved out in the
previous study. In the HCO$^+$ spectrum of the Turner-Welch object, a
weak, secondary velocity component at -44 km s$^{-1}$ is visible
(Fig. \ref{spectra}). This velocity component is not reported in any
previous studies. The HCN spectra towards the Turner-Welch object
shows an unexpected peak at an apparent velocity of -39 km s$^{-1}$
(Fig. \ref{spectra}); comparing relative positions of these two peaks,
the anomalous HCN peak would be the F=0-1 transition of a secondary
velocity component at -44 km s$^{-1}$. The other hyperfine lines from
this velocity component are being masked by the hyperfine lines from
the primary velocity component. This can only happen if the separation
between the hyperfine lines is approximately the separation between
the two velocity components. The C$^{18}$O line may also show this
secondary component, however, it cannot be confirmed due to the
relatively low SNR of the C$^{18}$O emission.
The C$^{18}$O map does not show any absorption towards the continuum
source, unlike HCN, HCO$^+$, and CN. This is consistent with
observations done by \citet{wink94}, in which they surmised that
the line excitation temperature of HCO$^+$ is significantly lower than
the kinetic temperature of 90 K; but the C$^{18}$O excitation
temperature is not low enough to produce absorption. However, our
observed absorption in other lines is significant and can yield
information about the regions around the continuum source. The HCO$^+$
spectra, seen in absorption (Figure \ref{spectra}), looks similar to
an inverse P-Cygni profile which could indicate an expanding
shell-like structure or a strong stellar wind. The multiple hyperfine
components of HCN make it difficult to tell which gas is causing the
absorption; however, it has the same central velocity and line width
as the HCO$^+$ absorption indicating that both features are being
generated from the same region. The CN absorption also appears to have
similar spectral features as HCN; however the red-shifted CN emission
feature is at a level of less than 1$\sigma$ of the noise level, and
cannot be considered as a positive detection.
Another notable feature of the continuum absorption lines is their
shape. There is a double trough shape seen in HCO$^+$, the strongest
hyperfine component of HCN, and in CN (these spectra are centered on
the strongest hyperfine component). There is also some evidence of
this feature in the other hyperfine components; however they are
weaker and, as a result, noisier. According to the results of
\citet{wink94}, within the UCHII region of W3(OH) there is an embedded
O7 star. It is very likely that W3(OH) is a region surrounding a
bright massive star embedded in a dense cloud with a strong solar wind
clearing out the region around the star. In looking at a velocity
moment map in HCO$^+$ (HCO$^+$ was chosen due to its lack of hyperfine
components and contamination), there appears to be an East-West
velocity gradient on the order of 3 km s$^{-1}$. Since W3(OH) is
unresolved in these maps, higher resolution data of W3(OH) are
required to quantify and describe this effect (see below).
\subsection {N$_2$H$^+$}
Figure \ref{n2h+channel} shows N$_2$H$^+$ channel maps. If N$_2$H$^+$
were tracing the same dense material as CN, we would expect to see
very similar emission (and absorption) spectra. Most notable is the
lack of an absorption feature towards the continuum source, or
alternatively, of any emission in N$_2$H$^+$ at the continuum
position. Also, the rotation feature in CN noted above is not seen in
N$_2$H$^+$. The N$_2$H$^+$ emission appears to be concentrated in the
south western region of the map, roughly coincident with one of the
velocity components of CN; however, this is where the similarity
ends. In the N$_2$H$^+$ spectra (Figure \ref{spectra}), only 3 peaks
appear when there are 7 hyperfine lines within our window. The central
peak, at about -47.5 km s$^{-1}$, contains 3 hyperfine components
which, having only 0.487 MHz (1.56 km s$^{-1}$) between them, are
blended together. The peak centered at -42.5 km s$^{-1}$ contains
another three hyperfine components whose transition frequencies all
lie within 0.432 MHz (1.39 km s$^{-1}$) and would also appear
blended. The last hyperfine component is seen at -56 km s$^{-1}$ and
is the only one that is not blended. Our N$_2$H$^+$ map contains some
similarities to the ammonia (NH$_3$) map presented by \citet{wilson93}
and \citet{tieftrunk98}. The major difference between the two is the
lack of N$_2$H$^+$ emission around the Turner-Welch object. The major
similarity comes from an east-west elongated clump present in the
northeast corner of both the N$_2$H$^+$ emission and the NH$_3$
emission. This stands out since none of the other molecules show any
sign of emission from this region. The closest emission is the north
lobe of the possible CN outflow. It is unclear if these regions are
related.
It is possible that the regions of CN emission much more closely
matches the regions of N$_2$H$^+$ than is readily apparent, even
though the spatial peaks do not coincide. Since CN is significantly
weaker than N$_2$H$^+$, it is possible that emission, particularly in
the lower west side of the map that is detectable in N$_2$H$^+$, is
significantly below the noise threshold in CN. To emphasize this fact,
which may not be noticeable in the individual channel maps, we
constructed a channel map with 20\% contours from all five detected
species (Fig. \ref{combined}). Most noticeable in the -48 km s$^{-1}$
and -47 km s$^{-1}$ panels, CN (with several other of the species)
does seem to trace much of the N$_2$H$^+$ emission. In the -51 km
s$^{-1}$ and -50 km s$^{-1}$ panels, it appears that the N$_2$H$^+$
emission borders on the regions with emission in the other species. If
we assume that the two molecules do trace the same density gas, it is
quite possible that some of these regions are undergoing a chemical
reaction that selectively annihilates N$_2$H$^+$.
\subsection{B-array Observations}
From the observations described above, several significant features
warranted even higher resolution observations. The continuum source
is seen in absorption in some species; however, it remains
unresolved. CARMA's B-array should be able to probe the region at
significantly higher resolution than in our previous data. This will
additionally allow us to study the velocity structure of gas near the
continuum source. A velocity map of HCO$^+$ shows an East-West
velocity gradient of $\sim$ 3 km s$^{-1}$. HCO$^+$ was chosen
specifically since it does not have multiple hyperfine components
which, due to blending, could pollute the velocity map, and shows
absorption towards the continuum source. In addition, with these data
we can see if we can resolve the individual cores of the Turner-Welch
object. Previous high resolution continuum maps have shown evidence of
multiple cores within the Turner-Welch object which we would expect to
also see in corresponding high resolution spectral line maps. We
previously choose to map the region with the C, D, and E arrays to
study large scale features; with B-array alone we are able to map
features that require higher spatial resolution. If we combined the
B-array data with the rest, we will produce a map with a synthesized
beam that is larger than the beam of the B-array data alone, and
larger than some of the smallest features, obscuring them. In
addition, the necessity to self-calibrate the data and issues with
gain calibration (discussed in detail below) make a separate analysis
more informative.
Self calibration results in several issues that need to be
overcome. W3(OH)'s physical size is 0.01 pc in diameter
\citep{kawamura98} which corresponds to an angular size of $\sim$
1\arcsec. This means that many of the longer baselines in B-array will
be resolving the W3(OH) continuum source. Data from these baselines
will make it difficult to self calibrate the data as it assumes a
point source model for the source (or requires an accurate model of
the source features which we do not have). Running self-cal on the
full dataset (including resolved baselines) generates a solution that
is appropriate for the data up until a UV-radius of 130k$\lambda$, at
which the solution fails. This UV-radius corresponds to a physical
size of:
\begin{eqnarray*}
\Theta & \sim & \frac{\lambda}{D} \\
& \sim & \frac{\lambda}{2 * \mathrm{UVr}} \\
& \sim & 0.79\arcsec
\end{eqnarray*}
Therefore running self-cal on the full dataset assuming a point source
will result in large errors in the selfcal solution due to long,
resolved baselines. If we ``cut'' the dataset so we only use the data
with UV-radius $<$ 130k$\lambda$ for use in calculating the phase
solution, we will not have this issue.
The selfcal technique fits the phase center of the data to the
brightest point source in the map. Our C, D, and E array data is
slightly offset in position from the pointing center, so this will
result in a slight position offset between the two maps. The can be
corrected for by manually fitting and entering the center position of
the continuum source from our C, D, E maps, however this is not
necessary since we will not be directly combining our B-array data
with the rest, since the resulting maps at full angular resolution
would have a very low signal-to-noise ratio.
This self-calibration technique also poses several problems with
regards to flux and amplitude calibration. Several dishes primarily
have long baselines, with only a few baselines that fall under our
UV-radius limit of 130k$\lambda$. While this worked well for phase
calibration, the results of amplitude selfcal was very poor. Many of
the amplitude gains, which should be flat and close to unity, were
noisy and had many datapoints corresponding to unphysical gains that
directly translates to poor image quality. The few remaining baselines
also tend to have low signal to noise which further increases the
problem.
In the attempt to apply correct gain calibration, several other
techniques were tested to amplitude calibrate the data. We attempted
to amplitude calibrate our data using the dedicated phase calibrator
measurements and transferring the resulting solution to our
source. This also failed, and while it produced better results than the
amplitude selfcal technique described above, the high resolution, low
intensity features were washed out in the final maps. This is most
likely due to the gain calibrator being 10$^o$ from our source, far
enough that the telescopes are viewing different enough atmosphere to
negatively affect the calculated gains. This problem does not affect
our C, D, E array data because the maximum baseline length is
significantly smaller in these arrays and are not as sensitive to
atmospheric fluctuations as the longer B-array baselines. Since these
several technique turned out poor results, it was decided not to use
amplitude (flux) calibration on the B-array data, which is not an
uncommon result when using selfcal. This is another reason why we do
not combine this data with our C, D, E data as we would need an
accurate gain solution to provide proper weights to combine the
datasets.
\subsection{Continuum}
Figure \ref{continuumall} represents the combined C, D, E array and
B-array only continuum maps. Each map was produced using
multi-frequency synthesis (MFS) on two 500 MHz windows, one in the
N$_2$H$^+$ band (average frequency of 91.2 GHz, not shown), and one in
the CN band (average frequency of 112 GHz, shown). In Figure
\ref{continuumall}a, neither the Turner-Welch object nor W3OH are
resolved. The peak flux of W3OH decreases by a factor of 2 from 91.2
GHz to 112 GHz, while the center of the Turner-Welch object increases
in flux by a factor of two over the same range. In the B-array only
data (Figure \ref{continuumall}b), we resolve two individual clumps
within the Turner-Welch object and slightly resolve structure within
W3OH (note the off-center position of the highest level contour in
Fig. \ref{continuumall}b). This structure supports the findings of
\citet{chen06}. We additionally see the same change in relative fluxes
in the B-array maps as we saw in the C, D, E array maps, over the same
frequency range.
\section{Discussion}
\subsection{W3(OH) Absorption Feature}
As stated above, W3(OH) itself is a UCHII region with an embedded O7
star. Our data show that several of the molecular tracers we have
observed can be seen in absorption towards this source. The nature and
structure of this colder absorbing gas may be able to give us some
insight into the properties of W3(OH). Figure \ref{hco+cuts} shows
three spectra sampled across the continuum source in HCO$^+$. HCO$^+$
was chosen for this since it does not have any hyperfine components
that could cause line profile confusion. C$^{18}$O, while also not
having any hyperfine components, was not chosen to trace this feature
as it is not seen in emission or absorption towards W3(OH). The
composite emission and absorption spectra are consistent with a dense,
hot region (containing HCO$^+$) surrounded by one or more cold,
less-dense layers which also contain HCO$^+$. The absorption feature
is evidence of cold, optically thick gas which is at a velocity of -46
km s$^{-1}$ and has a FWHM of 4 km s$^{-1}$. In looking at
non-continuum subtracted spectra (Figure \ref{hco+cuts}b), this
absorption line is saturated and extends completely down to zero
flux. This implies that this cold region is in front of the emission
region and the continuum source. The absorption feature is also seen
in CN and HCN, however, it is seen slightly in emission in
C$^{18}$O. The C$^{18}$O emission feature is also centered at -46 km
s$^{-1}$ and has a FWHM of 4 km s$^{-1}$, the same as the HCO$^+$
absorption. It is very likely that it is from the same region,
implying that the gas density is sufficient to raise the excitation
temperature of the C$^{18}$O line above that of the brightness
temperature of the continuum source so the line appears in emission,
while the higher critical densities of the HCO$^+$, HCN, and CN
transitions lead their excitation temperatures to be less than that
brightness temperature, so they are seen in absorption. To note:
Figure \ref{hco+cuts} is resolution limited since it is formed with C,
D, and E array data and its three positions only cover an area of
1.5-2 beamwidths. This was part of the motivation to acquire B-array
data as the same region would cover 5-8 beamwidths in a B-array map.
From the B-array spectra (Figure \ref{hcospectra}c), we only see
HCO$^+$ in absorption against the continuum emission. The lack of
emission either means that it is being resolved out at this high
spatial resolution, or that the beam is completely filled by continuum
emission, so we see no emission from the sides of the beam. In either
case, this means that the emission does not come from the same gas
that produces the absorption. To support this, we produced spectra
that only includes E-array (largest beam size, shortest baselines) in
order to inspect large scale emission. The E-array data (Figure
\ref{hcospectra}a) shows emission towards the continuum between -43 km
s$^{-1}$ and -50 km s$^{-1}$ with a self absorption feature at -46 km
s$^{-1}$. This high velocity gas is not seen in C$^{18}$O and
therefore seems to be hot, optically thin gas that is behind the
continuum source.
Therefore, we propose a multi-layered model of the region around
W3(OH). Behind the UCHII region is a high velocity, hot, large spatial
scale, optically thin region which appears to peak in intensity
towards the Turner-Welch object (and may very well be associated with
it). The UCHII region is optically thin, and has a small spatial scale
of about 0.7\arcsec in diameter which is extremely bright in the
continuum, enough such that the narrow band 2 MHz windows are
contaminated by continuum emission. In front of the continuum is a
very cold, optically thick region at a velocity of -46 km s$^{-1}$
which we primarily see in absorption, but we see slightly in emission
in C$^{18}$O.
\subsection{CARMA Flux}
Our maps are comprised of C, D, and E array data and do not cover the
entire UV space. We do not have zero-spacing data which would be
required in order to reconstruct the most accurate map of the
region. The IRAM data consists of only a single pointing and its beam
does not cover the full area of the CARMA maps. In order to quantify
the amount of flux that could be resolved out by using CARMA, we
examined the single dish data taken by \citet{falgarone08}. Since the
IRAM-30 meter telescope has a resolution of 23\arcsec, we smoothed our
CARMA maps to match by convolving the data with a 23\arcsec\ Gaussian
beam (Figure \ref{iramcomp}). Spectra, (\ref{iramcom2}) were then
extracted from the position corresponding to the IRAM-30 meter
pointing center, as well as two other positions centered on the
north-eastern and north-western components. According to the IRAM
technical documentation \citep{iram30m}, the main beam efficiency is
78\% for the band that includes CN. To compare the flux values, an
estimate of CARMA's main beam efficiency is needed
\citep{white08}. CARMA does not have published main beam efficiency
measurements, however, it can be estimated from single-dish aperture
efficiency measurements which vary from 55\% to 70\% depending on the
individual dish. From this, we estimate that the main beam efficiency
(per dish) would be $\sim$70-80\%. By adopting a value of 75\%,
approximately the same as IRAM's, we can directly compare the IRAM and
CARMA spectra. The IRAM peak was 2.2 K, while CARMA's was 1.4 K. This
means that the peak of the CARMA spectra was at 64\% of IRAM's (the
area of the CARMA spectra is 48\% of the IRAM spectra). CN is also
seen in strong absorption towards the central source (W3OH), however,
this absorption feature is not present in any of the IRAM data. This
is possible for several reasons. First, the relatively small spatial
size of the absorption feature covers a small fraction of the total
IRAM beam. Any measured absorption would be diluted across the entire
beam and masked by the significantly greater emission. In addition,
the IRAM data baseline fitting procedure could contribute to hiding
the presence of a weak absorption feature.
\subsection{N$_2$H$^+$ Chemical Reaction}
In regions with standard CO abundances (i.e. regions that are not
depleted in CO), one of the formation mechanisms of HCO$^+$ is the
destruction of N$_2$H$^+$ by CO \citep{bergin02,jorgensen04}. Since
there is significant support that CN remains in the gas phase at
densities greater than 10$^5$ cm$^{-3}$ (same as N$_2$H$^+$) we would
expect CN and N$_2$H$^+$ to have a very similar spatial distribution,
however, there are some regions and velocities where we would expect
to see more N$_2$H$^+$ emission than we do. The particular reaction
that is thought to occur is:
\begin{equation}
N_2H^+ + CO \rightarrow HCO^+ + N_2
\end{equation}
If we compare the regions of strong N$_2$H$^+$ and HCO$^+$ emission,
we would expect them not to be coincident. In addition, we expect CO
to be depleted in regions of strong N$_2$H$^+$ emission, as the
presence of CO would cause the formation of HCO$^+$, while regions
with significant CO emission must be weak in N$_2$H$^+$, or the
reaction between these two molecules would occur.
In the region surrounding the Turner-Welch object we can see
considerable emission from both CO and HCO$^+$, however, there is no
N$_2$H$^+$ within our detection limit. This depletion of N$_2$H$^+$
could be caused by the amount of C$^{18}$O present along this
sightline. Since there is significant CN emission from this region, we
would also expect to see N$_2$H$^+$. The presence of CN and the
relative strengths of HCO$^+$ and C$^{18}$O is significant evidence of
this reaction occurring and that N$_2$H$^+$ is being consumed by this
reaction.
Above, we showed that there was a cold, dark region of gas towards the
continuum source where HCO$^+$ is seen in strong absorption along
with CN and HCN. Discounting the contribution from the continuum
source, C$^{18}$O is seen slightly in emission and N$_2$H$^+$ is seen
slightly in emission and absorption. The presence of CN absorption
implies conditions where N$_2$H$^+$ should also be strongly
detectable. If we assume that the primary formation mechanism of
HCO$^+$ in this region is by the CO and N$_2$H$^+$ chemical reaction,
the relative depletion of these two chemical species in this cold dark
cloud can be understood.
In regions with conditions similar to that of W3(OH), CN may be a more
reliable tracer than N$_2$H$^+$ to directly sample regions at high
densities. However, since the strongest CN hyperfine transition has
significantly weaker line emission than the strongest hyperfine
transitions of N$_2$H$^+$, it would require significantly greater
telescope time to achieve a comparable signal-to-noise ratio. In our
case, these sets of maps show not only that this chemical reaction is
occurring, but is happening most strongly both along the sightlines
associated with an active star forming clump, and in a cold cloud in
front of an extremely bright UCHII region.
\subsection{Structure of the Region Surriounding W3(OH)}
\citet{wilson91} developed a model of the region surrounding W3(OH)
based on C$^{18}$O, C$^{34}$S, and methanol (among others). They
proposed a multi-layer model of low density molecular gas surrounding
a high density molecular core. This low density cooler region extends
in from of W3(OH) as well. Self-aborbtion in their CS data supports
this model. In addition, the $v_{LSR}$ of the self-absorbtion was up
to 2 km s$^{-1}$ more positive than the $v_{LSR}$ of the hotter
emitting region. This implys that the cloud envelope is contracting
relative to the cloud core. We see similar offsets in the E-array
spectrum of HCO$^{+}$ seen in Fig. \ref{hcospectra}a, however the
offset appears to decrease at higher angular resolutions
(Fig. \ref{hco+cuts}).
We can compare the range of radial velocities at which we see
molecular material. \citet{wilson91} data showed that C$^{18}$O peaks
towards W3(OH) at a radial velocity of -46.6 km s$^{-1}$, and peaks
towards the Turner-Welch object at -48 km s$^{-1}$. We see the same
radial velocities in our C$^{18}$O data (even though our dataset was
at 3mm, and theirs was at 1mm). For other species that trace related
density regimes (HCN and HCO$^{+}$), we see them peak at a velocity of
-47 km s$^{-1}$ towards the Turner-Welch object. These species are
seen strongly in absorbtion towards W3(OH), which makes it difficult
to determine its radial velocity. These data show that the region
around W3(OH) is extremely complex.
\section{Conclusions}
We mapped the 1 arcmin$^2$ region around W3(OH) with CARMA, in which
previous single-dish CN Zeeman observations detected a strong magnetic
field. In determining which gas was producing the magnetic field, we
were able to see the following features and make the following
conclusions from their presence:
\begin{enumerate}
\item A strong absorption feature in HCO$^+$, HCN, and CN was detected
towards the continuum source whose strength and line saturation
implies the existence of a cold, optically thick region in front of
the continuum source.
\item Comparison between CARMA spectra and spectra obtained with the
IRAM-30 meter telescope \citep{falgarone08} yields that CARMA
detects approximately 65\% of the material in this region.
\item Selective depletion of N$_2$H$^+$ in comparison with the CN and
HCO$^+$ distribution indicates that N$_2$H$^+$ is reacting with CO
to form HCO$^+$. This reduces the effectiveness of N$_2$H$^+$ as a
high-density tracer, and favors the use of species, such as CN, in
objects with similar chemical composition to W3(OH). These
measurments additionally support the usefulness of CN as a high
density tracer and confirms the hypothesis that CN Zeeman mapping
will probe the magnetic field strength in the high density regions
of molecular cloud clumps.
\end{enumerate}
These conclusions lead to the result that future CN Zeeman mapping at
high resolution with an interferometer is feasible and that W3(OH) is
an ideal target to perform such mapping on.
\section{Acknowledgments}
Support for CARMA construction was derived from the states of
California, Illinois, and Maryland, the James S. McDonnell Foundation,
the Gordon and Betty Moore Foundation, the Kenneth T. and Eileen
L. Norris Foundation, the University of Chicago, the Associates of the
California Institute of Technology, and the National Science
Foundation. Ongoing CARMA development and operations are supported by
the National Science Foundation under a cooperative agreement (NSF AST
08-38226), and by the CARMA partner universities.
|
\section{Introduction}
Let us start with seting up some notation and terminology. Assume $V$ is a finite dimensional normed
space. A subset $C$ is a {\em convex cone} (or simply {\em cone}) in $V$ if $\alpha x+\beta
y\in C$ for all $x,y\in C$ and $\alpha,\beta\in\bR_+$. A cone $V$ is {\em pointed} if $V\cap(-V)=\{0\}$.
A convex subcone $F\subset C$ is called a {\em face} if $x-y\in F$ implies $y\in F$ for every $y\in C$.
Any face $F$ such that $F\neq\{0\}$ and $F\neq C$ will be called {\em proper}.
A proper face $F$ is said to be {\em maximal} if for any face $G$ such that $F\subset G\subset C$ we have either
$G=F$ or $G=C$. If $K\subset C$ is any subset then by $F(K)$ we will denote the smallest face containing $K$.
If $K=\{x\}$ for some $x\in C$ then we will write $F(x)$ instead of $F(\{x\})$. An element $x\in C$ will be called
{\em extremal} if $F(x)=\overline{\bR_+}x$, where $\overline{\bR_+}$ is the set of all non-negative real numbers.
The set of all extremal elements will be denoted be $\mathrm{ext}\,C$. In the sequel we will need the following
\begin{lemma}\label{l:extF}
If $C$ is a cone and $F\subset C$ is a face then $\mathrm{ext}\,F=F\cap\mathrm{ext}\,C$.
\end{lemma}
\begin{proof}
Let $x\in\mathrm{ext}\,F$. We should show that $x\in\mathrm{ext}\,C$. Assume that $y\in C$ and $x-y\in C$. Since $x\in F$ and $F$ is a face, $y\in F$ and $x-y\in F$. Now, extremality of $x$ in $F$ implies $y=\lambda x$ for some nonnegative constant $\lambda$. Thus $x\in\mathrm{ext}\,C$. We proved $\mathrm{ext}\,F\subset F\cap\mathrm{ext}\,C$. The converse inclusion is obvious.
\end{proof}
Assume now that $W$ is another finite dimensional normed space, and $V$ and $W$ are dual to each other with respect to bilinear pairing $\langle\cdot,\cdot\rangle_{\mathrm{d}}$.
For a subset $C\subset V$ (respectively $D\subset W$) we define the {\em dual cone} $C^\circ=\{y\in W:\,\mbox{$\langle x,y\rangle_{\mathrm{d}}\geq 0$ for all $x\in C$}\}$ (respectively $D^\circ=\{x\in V:\,\mbox{$\langle x,y\rangle_{\mathrm{d}}\geq 0$ for all $y\in D$}\}$). One can show that $C^{\circ\circ}$ is the smallest closed cone containing $C$.
Assume $C\subset V$ is a closed convex cone and $F\subset C$ is a face. We define
$F'=\{y\in C^\circ:\,\mbox{$\langle x,y\rangle_{\mathrm{d}}=0$ for all $x\in F$}\}$.
It is clear that $F'$ is a closed face of $C^\circ$.
We say that a face $F$ of a closed convex cone $C$ is {\em exposed} if there exists $y_0\in C^\circ$
such that $F=\{x\in C:\,\langle x,y_0\rangle_{\mathrm{d}}=0\}$.
We have the following
\begin{lemma}[\cite{Kye00}]\label{l:Kye}
Let $F$ be a closed face of a closed convex cone $C$. Then $F$ is exposed if and only if
$F=F''$. The set $F''$ is the smallest closed face containing $F$.
\end{lemma}
An element $x\in C$ is called an {\em exposed point} of $C$ if $\overline{\bR_+}x$ is an exposed face
of $F$. The set of all exposed points of $C$ will be denoted by $\mathrm{exp}\,C$.
Now, let $\cK$ and $\cH$ be finite dimensional Hilbert spaces and $\dim\cK=m$, $\dim\cH=n$.
By $\fB(\cK)$ (respectively $\fB(\cH)$) we denote the $C^*$-algebra of all linear transformations
acting on $\cK$ (respectively $\cH$).
Let $V=\fB(\fB(\cK),\fB(\cH))$ be the space of linear mappings from $\fB(\cK)$ into $\fB(\cH)$
and let $W=\fB(\cH)\otimes\fB(\cK)$.
Assume also that some antilinear selfadjoint involutions $\cK\ni\xi\mapsto\overline{\xi}\in\cK$ and
$\cH\ni\eta\mapsto\overline{\eta}\in\cH$ are given.
Following \cite{Sto86} (see also \cite{Kye00}) we define the following bilinear pairing between $V$ and $W$
\begin{equation}
\langle \phi,X\otimes Y\rangle_{\mathrm{d}}=\Tr\left(\phi(Y)X^{\mathrm{T}}\right)
\end{equation}
where $\phi\in V$, $X\in\fB(\cH)$, $Y\in\fB(\cK)$ and $\mathrm{T}$ is the transposition on $\fB(\cH)$
determined by the given antilinear selfadjoint involution on $\cH$.
Now, we choose some special convex cones $\fP\subset V$ and $\fS\subset W$. Namely $\fP$ consists of positive maps,
i.e. such maps $\phi$ that $\phi\left(\fB(\cK)_+\right)\subset\fB(\cH)_+$, while $\fS=\fB(\cH)_+\otimes\fB(\cK)_+$
(its elements are sometimes called unnormalized separable states).
It is known that these cones are dual to each other, i.e. $\fS=\fP^\circ$ (e.g. \cite{MM01}).
The structure of extremal elements of $\fS$ is simple. For two vestors $x,y$ from a Hilbert space $\mathcal{X}$ we let $xy^*$ denote the rank $1$ operator on $X$ such that $(xy^*)z=\langle y,z\rangle x$ for $z\in\mathcal{X}$. It follows from the definition of $\fS$ that
$$\mathrm{ext}\,\fS=\{\xi\xi^*\otimes\eta\eta^*:\,\xi\in\cH,\,\eta\in\cK\}.$$
The big challenge is to describe the structure of extremal elements of the cone $\fP$. The full description
of the set $\mathrm{ext}\,\fP$ is still not done. Only some partial results are known. In \cite{Sto63} extremal elements in the convex set of unital maps acting from $\fB(\bC^2)$ into $\fB(\bC^2)$ are characterized.
As regards the full cone $\fP$ of positive but not necessarily unital maps, it was proved in
\cite{YH05} that every maps of the form
\begin{equation}\label{r:dec}
\phi(X)=AXA^*\quad\mbox{or}\quad\phi(X)=AX^{\mathrm{T}}A^*,\qquad X\in\fB(\cK),
\end{equation}
($A\in\fB(\cK,\cH)$) are extremal in $\fP$.
Moreover, several examples of non-decomposable extremal positive maps are scattered over the literature (see e.g. \cite{Cho75,Wor76,Rob85,Ha03,Bre06}).
Let us remind that due to Straszewicz's theorem (\cite{Str35}, see also \cite{Roc70}) the set $\mathrm{exp}\,\fP$ is dense in
$\mathrm{ext}\,\fP$.
Thus, in order to do a full characterization of positive maps it is enough to describe fully the set
of all exposed points of $\fP$. It was proved in \cite{YH05} that if $\rank A=1$ or $\rank A=m$ then
a map of the form (\ref{r:dec}) is an exposed point of $C$.
Recently, some new examples of exposed nondecomposable positive maps has appeared in the literature (see e.g. \cite{HK11,Chr11,ChS12a,SCh12,ChS12b}).
The aim of this paper is to provide some new examples of exposed positive maps. We will use results concerning rank properties of extremal positive maps described in \cite{Mar09}.
\section{Main result}
Now we are ready to formulate our main theorem.
\begin{theorem}\label{t}
Every map of the form (\ref{r:dec}) is an exposed point of $C$.
\end{theorem}
In the proof of the above theorem we will need more or less known two lemmas.
For any $A\in\fB(\cK,\cH)$ we denote $\Vert A\Vert_2=\left(\Tr(A^*A)\right)^{1/2}$.
\begin{lemma}\label{l:1to1}
There is a unique linear map $\fB(\cK,\cH)\ni A\mapsto f_A \in(\cH\otimes\cK)^*$
such that $f_A(\xi\otimes\eta)=\langle\overline{\xi},A\eta\rangle$ for any $\xi\in\cH$ and $\eta\in\cK$.
Moreover, we have $\Vert f_A\Vert=\Vert A\Vert_2$ for $A\in\fB(\cK,\cH)$.
\end{lemma}
\begin{proof}
Observe that $\cH\times \cK\ni(\xi,\eta)\mapsto\langle\overline{\xi},A\eta\rangle\in\bC$ is a bilinear
form. It follows from the universality property of a tensor product that this form has a unique
lift to a linear functional $f_A$ on the $\cH\otimes\cK$.
Now, if $f\in (\cH\otimes\cK)^*$ then we define $\varphi(\xi,\eta)=f(\overline{\xi}\otimes\eta)$
for $\xi\in\cH$ and $\eta\in\cK$. It is a sesquilinear form on $\cH\times\cK$, so there is
$A\in\fB(\cK,\cH)$ such that $\varphi(\xi,\eta)=\langle\xi,A\eta\rangle$. Hence we have
$f(\xi\otimes\eta)=\varphi(\overline{\xi},\eta)=\langle\overline{\xi},A\eta\rangle=f_A(\xi\otimes\eta)$.
It remains to show the equality of norms. Let $\eta_1,\ldots,\eta_m$ be an orthonormal basis of $\cK$ and
$\xi_1,\ldots,\xi_m$ be any system of vectors from $\cH$. Observe that
$\left\Vert\sum_i\xi_i\otimes\eta_i\right\Vert^2=\sum_i\Vert\xi_i\Vert^2$ and $\Vert A\Vert_2^2=\sum_i\Vert A\eta_i\Vert^2$.
Thus
\begin{eqnarray*}
\left|f_A\left(\sum_i\xi_i\otimes\eta_i\right)\right
&=&
\left|\sum_i\langle\overline{\xi_i},A\eta_i\rangle\right|\leq
\sum_i\left\Vert\overline{\xi_i}\right\Vert\Vert A\eta_i\Vert\leq \\
&\leq &
\left(\sum_i\Vert\xi_i\Vert^2\right)^{1/2}\left(\sum_i\Vert A\eta_i\Vert^2\right)^{1/2}=
\left\Vert\sum_i\xi_i\otimes\eta_i\right\Vert\Vert A\Vert_2.
\end{eqnarray*}
Hence $\Vert f_A\Vert\leq\Vert A\Vert_2$. Now, observe that
$\left\Vert\sum_i\overline{A\eta_i}\otimes\eta_i\right\Vert=\left(\sum_i\left\Vert A\eta_i\right\Vert^2\right)^{1/2}=\Vert A\Vert_2$ and
$\left|f_A\left(\sum_i\overline{A\eta_i}\otimes\eta_i\right)\right|=\Vert A\Vert_2^2=\Vert A\Vert_2\left\Vert\sum_i\overline{A\eta_i}\otimes\eta_i\right\Vert$. Thus we conclude $\Vert f_A\Vert=\Vert A\Vert_2$.
\end{proof}
\begin{lemma}\label{l:my}
Let $A\in\fB(\cK,\cH)$ and $\rank A\geq 2$. Assume that an operator $B\in\fB(\cK,\cH)$
satisfies the following condition: for any $\xi\in\cH$ and $\eta\in\cK$ if $\langle\xi,A\eta\rangle=0$ then
$\langle\xi,B\overline{\eta}\rangle=0$. Then $B=0$.
\end{lemma}
\begin{proof}
Let $\eta_1,\eta_2,\ldots,\eta_m$ be such that $\overline{\eta_1},\overline{\eta_2},\ldots,\overline{\eta_m}$ form
an orthonormal system of eigenvectors of the operator $A^*A$.
Let $r=\rank A$. Thus we may assume that $\overline{\eta_{r+1}},\ldots,\overline{\eta_m}$ correspond to zero
eigenvalue while
$\overline{\eta_1},\ldots,\overline{\eta_r}$ correspond to non-zero eigenvalues. Thus vectors $A\overline{\eta_1},\ldots,A\overline{\eta_r}$ span the image of the operator $A$.
Take any $j\in\{r+1,\ldots,m\}$.
Observe that $A\overline{\eta_j}=0$, so $\langle\xi,A\overline{\eta_j}\rangle=0$ for each $\xi\in\cH$.
It follows from the assumption of the
Lemma that $\langle\xi,B\eta_j\rangle=0$ for any $\xi\in\cH$, so $B\eta_j=0$.
Now, let $j\in\{1,\ldots,r\}$. It follows from the assumption that for any $\xi\in A\cK^\bot$ we have $\langle\xi,B\eta_j\rangle=0$. Hence $B\eta_j\in A\cK$.
Now consider also some $k\in\{1,\ldots,r\}$ such that $k\neq j$. For every $z\in\bC$ define
$$\zeta_z=-\overline{z}\Vert A\overline{\eta_k}\Vert^2 A\overline{\eta_j}+\Vert A\overline{\eta_j}\Vert^2 A\overline{\eta_k},\quad
\rho_z=\overline{\eta_j}+z\overline{\eta_k}.$$
Observe that $\langle\zeta_z,A\rho_z\rangle=0$. On the other hand we have
\begin{eqnarray*}
\langle\zeta_z,B\overline{\rho_z}\rangle
&=&-z\Vert A\overline{\eta_k}\Vert^2\langle A\overline{\eta_j},B\eta_j\rangle-|z|^2\Vert A\overline{\eta_k}\Vert^2\langle A\overline{\eta_j},B\eta_k\rangle+\\
&&+\Vert A\overline{\eta_j}\Vert^2\langle A\overline{\eta_k},B\eta_j\rangle+
\overline{z}\Vert A\eta_j\Vert^2\langle A\overline{\eta_k},B\eta_k\rangle.
\end{eqnarray*}
It follows from the assumption that this expression is equal to zero for every $z\in\bC$.
Thus we conclude that $\langle A\overline{\eta_k},B\eta_j\rangle=0$ for any $k=1,\ldots,r$.
Since $B\eta_j\in A\cK$ and $A\overline{\eta_1},\ldots,A\overline{\eta_r}$ span $A\cK$, we conclude
that $B\eta_j=0$.
Thus we proved that $B\eta_j=0$ for any $j=1,2,\ldots,m$. The Lemma follows from the fact that $\eta_1,\ldots,\eta_m$ is a basis of $\cK$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{t}]
Let us consider a map $\phi\in C$.
It follows from Lemma \ref{l:Kye} that $\phi$ is an exposed point if and only if $\{\phi\}''=\overline{\bR_+}\phi$.
Let us calculate the face $\{\phi\}'\subset D$ firstly. Since any closed convex cone in finite dimensional space $W$
is a closed convex hull of its extremal elements, in order to determine the face $\{\phi\}'$ it is enough to describe
its extremal elements. They are those elements of $\{\phi\}'$ which are extremal in $D$ (cf. Lemma \ref{l:extF}).
Thus, one should find all pairs $(\xi,\eta)\in\cH\times\cK$ such that $\xi\xi^*\otimes\eta\eta^*\in\{\phi\}'$.
An easy calculation shows that this holds if and only if
$\langle\overline{\xi},\phi(\eta\eta^*)\overline{\xi}\rangle=0$ or,
equivalently, $\phi(\eta\eta^*)\overline{\xi}=0$. As a consequence we get the following characterization
of the face $\{\phi\}''$: if $\psi\in C$, then $\psi\in\{\phi\}''$ if and only if $\psi(\eta\eta^*)\overline{\xi}=0$
for all pairs $(\xi,\eta)\in\cH\times\cK$ such that $\phi(\eta\eta^*)\overline{\xi}=0$.
Now, assume that $\phi(X)=AXA^*$ where $A$ is some linear map from $\cK$ into $\cH$.
One can easily show that $\xi\xi^*\otimes\eta\eta^*\in\{\phi\}'$ if and only if $\langle\overline{\xi},A\eta\rangle=0$. We will show that for $\psi\in C$
if $\psi\in\{\phi\}''$ then $\psi$ is rank 1 non-increasing in the sense of \cite{Mar09}.
Let $\eta\in\cK$. Consider any $\xi\in\cH$ such that
$\xi\bot A\eta$. Then $\overline{\xi}\overline{\xi}^*\otimes\eta\eta^*\in\{\phi\}'$ what is equivalent to
$\phi(\eta\eta^*)\xi=0$. So, it follows from the preceding paragraph that $\psi(\eta\eta^*)\xi=0$ for any $\xi\in\{A\eta\}^\bot$. Thus $\psi(\eta\eta^*)$ is a non-negative multiple of rank 1 positive element $(A\eta)(A\eta)^*$.
Now, it follows from Theorem 2.2 in \cite{Mar09} that we have three possibilities:
\begin{enumerate}
\item[(i)] $\psi(X)=\omega(X)Q$ for some positive functional $\omega$ on $\fB(\cK)$ and a $1$-dimensional projection $Q$ on $\cH$,
\item[(ii)] $\psi(X)=BXB^*$ for some $B\in\fB(\cK,\cH)$,
\item[(iii)] $\psi(X)=BX^{\mathrm{T}}B^*$ for some $B\in\fB(\cK,\cH)$.
\end{enumerate}
Assume firstly that $\psi$ satisfies the condition (ii), i.e. $\psi(X)=BXB^*$ for some $B\in\fB(\cK,\cH)$.
From the above considerations we conclude that $B$ satisfies the following condition: for any $(\xi,\eta)\in\cH\times\cK$, if $\langle\overline{\xi},A\eta\rangle=0$ then $\langle\overline{\xi},B\eta\rangle=0$.
Now apply Lemma \ref{l:1to1}. Using notations introduced there we can write the above condition as
$\ker f_A\subset\ker f_B$. It is equivalent to the fact that $f_B$ is a multiple of $f_A$. Hence $B=\lambda A$
for some $\lambda\in\bC$, and $\psi=|\lambda|^2\phi$, and consequently $\psi\in\overline{\bR_+}\phi$.
Secondly, consider the case (i), i.e. let $\psi(X)=\omega(X)Q$, where $\omega$ is some positive functional on $\fB(\cK)$
and $Q$ is a $1$-dimensional projection on $\cH$.
Let $\omega(X)=\Tr(RX)$ where $R$ is some positive operator on $\cK$ and let $Q=\zeta\zeta^*$ for some non-zero $\zeta\in\cH$.
We will show that $\rank R\leq 1$.
To this end observe that the condition $\psi\in\{\phi\}''$ is equivalent to the following:
for any $(\xi,\eta)\in\cH\times\cK$, if $\langle\xi,A\eta\rangle=0$ then $\langle\eta,R\eta\rangle=0$ or $\langle\zeta,\xi\rangle=0$.
It follows that $\langle\eta,R\eta\rangle=0$ provided that $A\eta=0$, hence $\ker A\subset\ker R$.
Assume that there are two vectors $\eta_1,\eta_2\in\cK$ such that $R\eta_1,R\eta_2$ are linearly independent.
Then $A\eta_1,A\eta_2$ are also linearly independent.
Fix $i\in\{1,2\}$. Since $\psi\in\{\phi\}''$,
for any $\xi\in\{A\eta_i\}^\bot$ we have $\langle\eta_i,R\eta_i\rangle=0$ or $\langle\zeta,\xi\rangle\zeta=0$. We assumed $\langle\eta_i,R\eta_i\rangle\neq 0$, so we conclude that $\zeta\bot\xi$. Thus we proved
that $\zeta\in\bC A\eta_i$ for $i=1,2$. But $A\eta_1,A\eta_2$ are independent, so $\zeta=0$ which is a contradiction
to the assumption. Thus we proved that $R=\rho\rho^*$ for some $\rho\in\cK$. Now, we can write
$\phi(X)=\zeta\rho^*X\rho\zeta^*=\zeta\rho^*X(\zeta\rho^*)^*$ and we arrived at the previously described case (ii).
Finally, assume that $\phi(X)=CX^{\mathrm{T}}C^*$ for some $C\in\fB(\cK,\cH)$.
Observe that in this case the condition $\psi(\eta\eta^*)\overline{\xi}=0$ is equivalent to $\langle\overline{\xi},C\overline{\eta}\rangle=0$. Thus $\psi\in\{\phi\}''$ if and only if
for any $(\xi,\eta)\in\cH\times\cK$, $\langle\overline{\xi},A\eta\rangle=0$ implies $\langle\overline{\xi},C\overline{\eta}\rangle=0$. It follows from Lemma \ref{l:my}
that if $\rank A\geq 2$ then $C=0$. It remains to consider the case when $\rank A\leq 1$. If $A\eta=0$ then $C\overline{\eta}=0$, hence $\overline{\ker A}\subset\ker C$, and consequently $\rank C\leq 1$.
Then $C=\zeta\rho^*$ for some $\zeta\in\cH$ and $\rho\in\cK$. Hence $\psi(X)=\langle\rho,X^{\mathrm{T}}\rho\rangle\zeta\zeta^*$. Since $X\mapsto\langle\rho,X^{\mathrm{T}}\rho\rangle$
is a positive functional on $\fB(\cK)$, we came to the case (i).
This ends the proof for $\phi(X)=AXA^*$.
If $\phi(X)=AX^{\mathrm{T}}A^*$ then the proof is similar. To show that any $\psi\in\{\phi\}''$ is a multiple
of $\phi$ one should firstly show that $\psi$ is rank 1 non-increasing, then consider cases (iii), (i) and (ii).
\end{proof}
|
\section{Problem}
The problem of the spatial resolution improvement without loosing spectral resolution
is one of the basic technical problems of the incoherent scatter (IS) technique (as well
as other remote sensing techniques, based on weak single backscattering from random media).
In such techniques the basic parameters that provide information about media properties are
spectral power, correlation function and correlation matrix of the scattered signal (see [1]).
One can show that first two functions can be easily obtained from the third one. From scattering theory
it is also known that correlation matrix of the received signal is an average of the irregularities
correlation function over the altitudes. The integration over altitudes is made with the weight defined by
sounding signal shape and processing technique, and statistical averaging $<...>$ corresponds to the
averaging over the sounding runs, see [1]:
\begin{equation}
<u(T)\overline{u}(T+\tau)>=\int<W(r-\frac{cT}{2},\tau)P(r)\sigma(r,\tau)>dr;\, W(r,\tau)=a(\frac{rc}{2})\overline{a}(\frac{rc}{2}+\tau)\label{eq:1}\end{equation}
In the literature the integration weight $W(r,\tau)$ is referred as 'ambiguity function'.
The power profile value $P(r)$ takes into account the radiowaves propagation, antenna pattern and
cross-section dependence on radar range, $\sigma(r,\tau)$ - correlation characteristics of the
scatterers and $a(t)$ - sounding signal shape.
So, the practical problem of the spatial resolution improvement (formulated as obtaining information
about correlation characteristics of the scatterers $\sigma(r,\tau)$ with integration over the smallest
distances range) mathematically leads to the problem of the sounding and processing techniques synthesis
that gives the ambiguity function $W(r,\tau)$ with necessary characteristics.
The problem is currently under investigation (see [2]). But, for now, there are two most useful techniques
for spatial resolution improvement: with using static amplitude-modulated signals (see [3]) and with
phase-modulated signals, that change theirs shape from one sounding to another (see [4]).
Constructing of these techniques is valid in the approximation of quasistationary media.
In this approximation one can neglect changes of the irregularities correlation characteristics during averaging time
and averaging over the sounding runs becomes equivalent to the statistical averaging. The use of such techniques allows,
from one side, to obtain high spatial resolution, defined by discrete of temporal variation of the sounding signal, and
from the other side to obtain correlation characteristics of the irregularities at large delays (lags), defined by
duration of the whole signal.
The well known difficulty of these techniques is impossibility to obtain the correlation function at all the lags with
the same high spatial resolution and signal-to-noise ratio. This is caused by the following peculiarities of these
techniques: correlation function of the irregularities at zero lag can not be obtained with high spatial resolution;
correlation function can not be obtained at all the real lags within pulse duration, but only at discrete number of lags;
the number of the 'correct' correlation function lags and spatial resolution are usually connected: the higher spatial
resolution - the more lags in correlation function you get. So one can not obtian a lot of lags, necessary for correct
estimation of the correlation function, when one need to obtain average spatial resolution (about 1/4 - 1/5 of the
whole pulse length).
We present sounding technique that allows to improve spatial resolution of incoherent scatter technique without the
problems described above.
\section{Solution}
From (\ref{eq:1}) one can see that independence of irregularities correlation function on sounding signal shape allows
one to use any linear combination of the correlation matrixes obtained during experiment instead of simple summarizing
over the sounding runs(i.e. averaging over the sounding runs). In this case the ambiguity function will be
(with accuracy of the custom multiplier) the same linear combination of the ambiguity functions for different sounding runs,
summarized over the sounding run number $j$ (the fact, for example, is used in summation rules of the alternating code
technique, see [4]):
\begin{equation}
\sum D_{j}u_{j}(T)\overline{u}_{j}(T+\tau)=\int<W(r-\frac{cT}{2},\tau)><P(r)\sigma(r,\tau)>dr;
\label{eq:2}\end{equation}
\begin{equation}
<W(r,\tau)>=\sum D_{j}a_{j}(\frac{rc}{2})\overline{a}_{j}(\frac{rc}{2}+\tau) \\
\label{eq:3}\end{equation}
The ideal ambiguity function must allow to determine irregularities correlation function without distortion effects, caused
by power profile $P(r)$. At Irkutsk IS radar power profile is very irregular due to both electron density profile and Faraday
variations profile (see [5]). One can show that the ideal ambiguity function must have the shape:
\begin{equation}
<W(r,\tau)>=A(r)B(\tau)\label{eq:4}\end{equation}
We have found that for obtaining the almost ideal ambiguity function we should use a simple rectangle pulse as a sounding signal,
with constant amplitude and with two custom durations, the pulse duration should change from one sounding to another:
$dom (a_{2n})=[0,T_{long}];\, dom (a_{2n+1})=[0,T_{short}]$;
during processing we should subtract the correlation matrix for shorter pulse from the correlation matrix for longer pulse.
In this case $D_{2n}=1;D_{2n+1}=-1$ in (\ref{eq:2}).
During the subtraction some parts of the ambiguity functions are effectively compensate each other, so we call this technique
'effective subtraction technique' (EST).
It is important to note that (for accurate positioning the ambiguity function) the correlation matrix must be calculated in the form:
\begin{equation}
R_{j}(T,\tau)=\left\{ \begin{array}{c}
u_{j}(T-\tau)\overline{u}_{j}(T)\, for\,\tau\geq0\\
u_{j}(T)\overline{u}_{j}(T+\tau)\, for\,\tau<0\end{array}\right\} \label{eq:5}\end{equation}
During the experiment, we should calculate correlation matrix (\ref{eq:5}), but separately for sounding by the longer pulse
$R_{long,2n}(T,\tau)$, and by shorter pulse $R_{short,2n+1}(T,\tau)$. The correlation matrix $R_{EST}(T,\tau)$ that provides
ambiguity function with the almost ideal shape (\ref{eq:4}) is calculated as combination of $R_{long,2n}(T,\tau)$ and $R_{short,2n+1}(T,\tau)$:
\begin{equation}
<R_{EST}(T,\tau)>=\sum R_{long,2n}(T,\tau)-\sum R_{short,2n+1}(T,\tau)\label{eq:5a}\end{equation}
The delay $T$ corresponding to the radar range should be measured from the sounding pulse start.
Ambiguity functions of the correlation matrix (\ref{eq:5}) for longer pulse, shorter pulse and the difference between them
(\ref{eq:5a}) are shown at Fig.1. From the Fig.1 one can see that ambiguity function of correlation matrix (\ref{eq:5a})
is close to the ideal one (\ref{eq:4}). One can show that the size of ambiguity function over the range (i.e. spatial resolution)
is proportional to the difference of the pulses duration.
Also, one can see, that the size of the ambiguity function over the lags (i.e. lag resolution) is proportional to the half-sum
of the pulses durations.
\begin{figure}
\label{fig1}
\includegraphics[scale=0.17]{fig1}
\caption{Ambiguity function for EST (\ref{eq:5}-\ref{eq:5a}): for correlation matrix for longer pulse sounding (A),for
correlation matrix for shorter pulse sounding(B), and for theirs difference (C). Range and lag are in longer pulse
duration units.
}
\end{figure}
\section{Experimental results}
The EST was tested at Irkutsk IS radar [5] during the experiment 22-23/12/2010. As a sounding pulse we used a simple
rectangle signal. Its duration was changed during sounding runs and was 900mks for longer pulse and 675mks for shorter pulse.
The amplitude of the signal remains constant. Using received signal, we calculated correlation matrix (\ref{eq:5}) separately
for longer pulse sounding runs $R_{long,2n}(T,\tau)$ and for shorter pulse $R_{short,2n+1}(T,\tau)$ sounding runs.
Correlation matrix that provides high spatial resolution was calculated as theirs combination (\ref{eq:5a}).
For direct estimation of the ambiguity function we have used sounding runs with observations of the localized
space scatterers. It can be shown that correlation matrix (\ref{eq:2}) in this case becomes close to the ambiguity function:
\begin{equation}
\sigma(r,\tau)=\delta(r-R_0);
<R_{EST}(T,\tau)>=<P(R_0)><W(R_{0}-\frac{cT}{2},\tau)>
\end{equation}
The correlation matrix for localized scatterer (i.e. ambiguity function shape) obtained experimentally for EST is
shown at Fig.2. As one can see, new technique actually improves the spatial resolution, as expected theoretically.
\begin{figure}[t]
\label{fig2}
\begin{center}
\includegraphics[scale=0.54]{FIG2.jpg}
\end{center}
\caption{The EST correlation matrix (\ref{eq:5}-\ref{eq:5a}) for scattering from localized space scatterer during experiment 23/12/2010.}
\end{figure}
At Fig.3 we present altitude dynamics of the spectral power for daytime and nighttime data, obtained with the technique
described and with standard technique based only on longer pulse data. From the data presented one can see that spectral
power valley near zero frequency for the new technique becomes deeper and can be observed even at the night data.
It should be noted that such a spectral power shape actually theoretically expected for the ionospheric conditions
(see [1]), so the new technique not only improves spatial resolution in comparison with standard technique, but slightly
improves spectral resolution too.
\begin{figure}[t]
\label{fig3}
\begin{center}
\includegraphics[scale=0.23]{fig3b}
\end{center}
\caption{Spectral power obtained with use of EST (dashed line), and standard Irkutsk IS radar technique (solid line) over the same data,
obtained during experiment. (A,C,E,G) – daytime data, (B,D,F,H) – nighttime data. Altitudes: (G,H) - 230km, (E,F) - 260km, (C,D) - 290km, (A,B) - 320km}
\end{figure}
\section{Conclusion}
In conclusion we summarize some properties of the new effective subtraction technique (\ref{eq:5}-\ref{eq:5a}):
the spatial resolution is equal for all the lags including zero lag, and correlation matrix can be calculated at
any lag with almost equal signal-to-noise ratio; the EST uses only long enough simple pulses without any modulation, so in the
case of the low signal-to-noise ratio (for example at high altitudes) we can use standard signal processing technique
for both pulses separately with standard spatial resolution, and for high signal-to-noise ratio (at low altitudes) -
use EST with high spatial resolution; the spatial resolution of EST is defined by pulse duration difference, and
maximal allowed lags are defined by half-sum of the pulses durations.
\section{References}
1. Hagfors T. ``Plasma Fluctuations Excited by Charged Particle Motion and their
Detection by Weak Scattering of Radio Waves" in D.Alcayde (ed.),
\emph{Incoherent Scatter Theory, Practice and Science, Collection of lectures
given in Cargese, Corsica, 1995. EISCAT Technical Report 97/53}, 1997, pp. 1-32.
2. M.S.Lehtinen , I.I.Virtanen, J.Vierinen, ``Fast comparison of IS radar code
sequences for lag profile inversion" \emph{Ann.Geophysicae}, 26, 2008,
pp. 2291-2301.
3. M.P.Sulzer, ``A radar technique for high range resolution incoherent scatter
autocorrelation function measurements utilizing the full average power of
klystron radars," \emph{Radio Sci}, 21(6), 1986, pp. 1033-1040.
4. A.Huuskonen, M.S.Lehtinen, J.Pirttil , ``Fractional lags in alternating codes:
Improving incoherent scatter measurements by using lag estimates at noninteger
multiples of baud length," \emph{Radio Sci.}, 32(6), 1996, pp. 2271-2282.
5. G.A.Zherebtsov, A.V.Zavorin, A.V.Medvedev , V.E.Nosov, A.P.Potekhin,
B.G.Shpynev, ``Irkutsk Incoherent Scatter Radar," \emph{Radiotekh. Elektron.
(Moscow)}, 47(11), 2002, pp. 1339–1345.
\end{document} |
\section{Introduction}
Many problems in information theory, communications, statistical signal
processing, and related disciplines can be formalized as being
about the quest for a strategy $s$ that
minimizes (or maximizes) the expectation
of a certain cost function, $\ell(X,s)$, where
$X$ is a random variable (or a random vector).
Just a few examples of this generic paradigm are the following: (i)
Lossless and lossy data compression, where $X$ symbolizes the data to be
compressed, $s$ is the data compression scheme, and $\ell(X,s)$ is the length
of the compressed binary representation, or the distortion (in the lossy
case) or a linear combination of both (see, e.g.,
\cite[Chapters 5 and 10]{CT06}).
(ii) Gambling and portfolio theory \cite[Chapters 6 and 16]{CT06}, where cost
function is logarithm of the wealth relative.
(iii) Lossy joint source--channel coding,
where $X$ collectively symbolizes the randomness of
source and the channel, $s$ is the encoding--decoding scheme and $\ell(X,s)$
is the distortion in the reconstruction (see, e.g.,
\cite{Wyner80},\cite{Wyner81}). (iv)
Bayesian estimation of a random variable based on measurements, where $X$
designates jointly the desired random variable and the measurements, $s$ is
the estimation function and $\ell(X,s)$ is the error function, for the
example, the squared error. Non--Bayesian estimation problems can be
considered similarly (see, e.g., \cite{VanTrees68}). (v)
Prediction, sequential decision problems (see, for example,
\cite{MF98}) and stochastic control problems \cite{Bertsekas07}, such as the linear
quadratic Gaussian (LQG) problem, as well as general Markov decision
processes, are also
formalized in terms of selecting strategies in order to minimize
the expectation of a certain loss function.
While the criterion of minimizing the expected value of $\ell(X,s)$ has been
predominantly the most common one, the exponential moments of $\ell(X,s)$,
namely, $\bE\exp\{\alpha\ell(X,s)\}$ ($\alpha > 0$), have received much less
attention than they probably deserve in this context.
There are a few motivations for
examining strategies that minimize exponential moments. First,
$\bE\exp\{\alpha\ell(X,s)\}$, as a function of $\alpha$, is obviously the
moment--generating function of $\ell(X,s)$, and as such, it provides the full
information about the entire distribution of this random variable, not just its first
order moment. Thus, in particular, if we are fortunate enough to find a strategy that
uniformly minimizes $\bE
\exp\{\alpha\ell(X,s)\}$ for all $\alpha \ge 0$ (and there are examples
that this may be the case), then this is much stronger than just minimizing the
first moment. Secondly, exponential moments are intimately related to
large--deviations rate functions, and so, the minimization of exponential
moments may give us an edge on minimizing probabilities of (undesired) large deviations
events of the form $\mbox{Pr}\{\ell(X,s) \ge L_0\}$ (for some threshold $L_0$), or more precisely,
on maximizing the exponential rate of decay of these probabilities. There are
several works along this line, especially in contexts related to
buffer overflow in data
compression \cite{FM86},\cite{Humblet81},
\cite{Jelinek68},\cite{Merhav91},\cite{MN92},\cite{UH99},\cite{Wyner74-1},
and exponential moments related to guessing \cite{Arikan96},\cite{AM98a},
\cite{AM98b},\cite{Massey94},\cite{MA99},\cite{MRA99}.
It is natural to ask, in view of the foregoing discussion, how can we harness
the existing body of knowledge concerning optimization of strategies for
minimizing the first moment of $\ell(X,s)$, which is quite mature in many
applications, in our quest for optimum strategies that minimize exponential
moments. Our main basic result, in this paper, is a simple theorem that
relates the two criteria. In particular, we
furnish sufficient conditions that
the optimum strategy in the exponential moment sense can be
found in terms of the optimum strategy in the first moment sense, for a
possibly different probability distribution, which our theorem characterizes.
In some applications, these
sufficient conditions for optimality
in the exponential moment sense, yield an equation in $s$, whose solution is
the desired optimum strategy. It is clear then that in these applications,
the optimality conditions provide a concrete tool for deriving the optimum
solution. In other applications, however, this may not be quite the case
directly, yet the set of optimality conditions may still
serve as a useful tool: More often than not, in a given instance of the problem
under discussion, one may have a natural intuitive
guess concerning the optimum strategy, and then the optimality conditions
can be used to prove
that this is the case.
One example for this, that will be demonstrated in detail later on, is the following:
Given $n$ independent and identically distributed
(i.i.d.) Gaussian observations, $X_1,\ldots,X_n$, with mean
$\theta$, the sample mean, $s(X_1,\ldots,X_n)=\frac{1}{n}\sum_{i=1}^nX_i$,
is the optimum unbiased estimator of $\theta$, not merely in the
mean squared error sense (as is well known), but also in the sense of minimizing all
exponential moments of the squared error, i.e.,
$\bE\exp\{\alpha[s(X_1,\ldots,X_n)-\theta]^2\}$ for all $\alpha \ge 0$ for which
this expectation is finite.
We next devote some attention to the asymptotic regime. Consider the case where
$X$ is a random vector of dimension $n$, $X=(X_1,\ldots,X_n)$, governed by
a product--form probability distribution, and $\ell(X,s)$ grows linearly for a
given empirical distribution of $X$, for example, when
$\ell(X,s)$ is additive, i.e., $\ell(X,s)=\sum_{i=1}^n l(X_i,s)$. In this case,
the exponential moments of $\ell(X,s)$ typically behave (at
least asymptotically) like exponential functions of $n$.
If we can then select a strategy $s$ that somehow ``adapts''\footnote{
The precise meaning of this will be
clarified in the sequel.}
to the empirical
distribution of $(X_1,\ldots,X_n)$, then such strategies may be universally
optimum (or asymptotically optimum in the sense of achieving the minimum
exponential rate of the exponential moment) in that they depend on neither the underlying probability
distribution, nor on the parameter $\alpha$. This is demonstrated in several
examples, one of which is an extension of a well known result by Rissanen in
universal data compression \cite{Rissanen84}.
An interesting byproduct of the use of the exponential moment criterion in
the asymptotic regime is
the possible existence of phase transitions: In turns out that the
asymptotic exponential rate of $\bE\exp\{\alpha\ell(X_1,\ldots,X_n,s)\}$ as a
function of $n$, may not be a smooth function of $\alpha$ and/or the parameters
of the underlying probability distribution even when the model under
discussion seems rather simple and `innocent.'
This is best understood from a
statistical--mechanical perspective, because in some cases, the calculation of
the exponential moment is clearly analogous to that of the partition function
of a certain physical system of interacting particles, which is known to
exhibit phase transitions. It is demonstrated that at least in certain cases, these phase
transitions are not merely an artifact of a badly chosen strategy, but they
appear even when the optimum strategy is used, and hence these phase
transitions are inherent in the model.
We end this paper by touching upon yet another aspect of the exponential
moment criterion, which we do not investigate
very thoroughly here, but we believe it is interesting and therefore
certainly deserves a further study in the future:
Even in the ordinary setting, of seeking strategies that minimize
$\bE\{\ell(X,s)\}$, optimum strategies may not always be known, and then
lower bounds are of considerable importance as a reference performance figure.
This is {\it a--fortiori} the case when exponential moments are considered. One
way to obtain non--trivial
bounds on exponential moments is via lower bounds the expectation of
$\ell(X,s)$, using the techniques developed in this paper.
We demonstrate this idea in the context of a lower bound on
the expected
exponentiated squared error of an unbiased
parameter estimator, on the basis of the
Cram\'er--Rao bound (CRB),
but it should be understood that, more generally, the same idea can be applied
on the basis of
other well--known bounds of the mean-square error (Bayesian and non--Bayesian)
in parameter estimation, and in
signal estimation,
as well as in other problem areas.
\section{Basic Optimality Conditions}
Let $X$ be a random variable taking on values in a certain alphabet ${\cal X}$,
and drawn according to a given probability distribution $P$. Let the variable
$s$ designate a {\it
strategy} chosen from some space ${\cal S}$ of allowed strategies. The term
``strategy'' in our context is fairly generic: it may be
a scalar variable, a vector, an infinite sequence, a function (of $X$), a
partition of ${\cal X}$, a coding scheme for $X$, and so on.
Associated with each $x\in{\cal X}$ and $s\in{\cal S}$, is
a loss $\ell(x,s)$. The function $\ell(x,s)$ is called the {\it loss function},
or the {\it cost function}.
The operator $\bE\{\cdot\}$ will
be understood as the expectation operator with respect to (w.r.t.) the
underlying distribution $P$, and whenever we refer to the expectation w.r.t.\
another probability distribution, say, $Q$, we use the notation
$\bE_Q\{\cdot\}$. Nonetheless, occasionally, when there is more than one probability
distribution playing a role at the same time and we wish to emphasize that the expectation is
taken w.r.t.\ $P$, then to avoid confusion, we may denote this expectation
by $\bE_P\{\cdot\}$.
For a given $\alpha > 0$, consider the problem of minimizing
$\bE\exp\{\alpha\ell(X,s)\}$ across $s\in{\cal S}$. The following theorem
relates the optimum $s$ for this problem to the optimum $s$ for the problem
of minimizing $\bE_Q\{\ell(X,s)\}$ w.r.t.\ another probability distribution
$Q$.
\begin{theorem}
Assume that there exists a strategy $s\in{\cal S}$ for which
\begin{equation}
Z(s)\stackrel{\mbox{(d)}}{=}\bE_P\exp\{\alpha\ell(X,s)\} < \infty.
\end{equation}
A strategy $s\in{\cal S}$
minimizes $\bE_P\exp\{\alpha\ell(X,s)\}$ if there exists
a probability distribution $Q$ on ${\cal X}$ that satisfies the following two
conditions at the same time:
\begin{enumerate}
\item The strategy $s$ minimizes $\bE_Q\{\ell(X,s)\}$ over ${\cal S}$.
\item The probability distribution $Q$ is given by
\begin{equation}
Q(x)=\frac{P(x)e^{\alpha\ell(x,s)}}{Z(s)}.
\end{equation}
\end{enumerate}
\end{theorem}
An equivalent formulation of Theorem 1 is the following: denoting
by $s_Q$ a strategy that minimizes $\bE_Q\{\ell(X,s)\}$ over ${\cal S}$,
then the theorem asserts that $s_Q$ minimizes $\bE_P\exp\{\alpha\ell(X,s)\}$
over ${\cal S}$ if
\begin{equation}
\label{cond}
Q(x) \propto P(x)e^{\alpha\ell(x,s_Q)},
\end{equation}
where by $A(x)\propto B(x)$, we mean that $A(x)/B(x)$ is a constant,
independent of $x$.
\vspace{0.2cm}
\noindent
{\it Proof.}
Let $s\in{\cal S}$ be arbitrary and let $(s^*,Q^*)$ satisfy conditions 1 and 2 of
Theorem 1. Consider the following chain of inequalities:
\begin{eqnarray}
\label{chain}
\bE_P\exp\{\alpha\ell(X,s)\} &=&
\bE_{Q^*}\exp\left\{\alpha\ell(X,s)+\ln\frac{P(X)}{Q^*(X)}\right\}\nonumber\\
&\ge&\exp\left\{\alpha\bE_{Q^*}\ell(X,s)-D(Q^*\|P)\right\}\nonumber\\
&\ge&\exp\left\{\alpha\bE_{Q^*}\ell(X,s^*)-D(Q^*\|P)\right\}\nonumber\\
&=&\exp\left\{\alpha\bE_{Q^*}\ell(X,s^*)-
\bE_{Q^*}\ln\frac{e^{\alpha\ell(X,s^*)}}{Z(s^*)}\right\}\nonumber\\
&=& Z(s^*) =
\bE_P\exp\{\alpha\ell(X,s^*)\},
\end{eqnarray}
where the first equality results from a change of measure (multiplying
and dividing $e^{\alpha\ell(X,s)}$ by $Q^*(X)$), the second line is by
Jensen's inequality and the convexity of the exponential function (with
$D(Q\|P)\stackrel{\mbox{(d)}}{=}\bE_Q\ln[Q(X)/P(X)]$ being the relative entropy between $Q$ and
$P$), the third line is by condition 1 of Theorem 1, and the remaining
equalities result from condition 2: On substituting
$Q^*(x)=P(x)e^{\alpha\ell(x,s^*)}/Z(s^*)$ into $D(Q^*\|P)$, one readily
obtains $D(Q^*\|P)=\alpha\bE_{Q^*}\ell(X,s^*)-\ln Z(s^*)$.
This completes the proof of Theorem 1.
$\Box$
Observe that for a given $s$, Jensen's inequality in the second line
of (\ref{chain}), becomes an equality for $Q(x)=P(x)e^{\alpha\ell(x,s)}/Z(s)$,
since for this choice of $Q$, the random variable that appears in the
exponent, $\alpha\ell(X,s)+\ln\frac{P(X)}{Q(X)}$, becomes degenerate
(constant with probability one). Since the original expression is independent
of $Q$, such an equality in Jensen's inequality means that
$\alpha\bE_{Q}\ell(X,s)-D(Q\|P)$ is maximized by this choice of $Q$, a fact
which can also be seen from a direct maximization of this expression using
standard methods.
Thus, we have a simple identity for every $s$:
\begin{equation}
\label{identity}
\bE_P\exp\{\alpha\ell(X,s)\}=\exp\{\alpha\max_Q[\bE_Q\ell(X,s)-D(Q\|P)]\}.
\end{equation}
This identity will prove useful in several places throughout the sequel.
Suppose next that the set ${\cal S}$ and the loss function $\ell(x,s)$
are such that:
\begin{equation}
\label{minimax}
\min_{s\in{\cal S}}\max_Q[\alpha\bE_Q\ell(X,s)-D(Q\|P)]=
\max_Q\min_{s\in{\cal S}}[\alpha\bE_Q\ell(X,s)-D(Q\|P)].
\end{equation}
This equality between the min--max and the max--min means that there is a
saddle point $(s^*,Q^*)$, where $s^*$ is a solution of the min--max
problem on the left--hand side and $Q^*$ is a solution to the max--min problem
on the right--hand side. It is easy to check
that the maximizing $Q$ in the inner
maximization on the left--hand side is
$Q^*(x)=P(x)e^{\alpha\ell(x,s^*)}/Z(s^*)$, which is
condition 2 of Theorem 1. By the same token,
the inner minimization over $s$ on the right--hand side obviously
minimizes $\bE_{Q^*}\ell(X,s)$, which is condition 1. It follows then
that if the min--max and the max--min are equal, then the saddle point
satisfies the conditions of Theorem 1, and hence the corresponding
$s^*$ is optimum. Note also that when eq.\ (\ref{minimax}) holds, the
conditions of Theorem 1 become also necessary conditions for optimality:
Suppose that $s^*$ is optimum. Then, by eq.\ (\ref{identity}), it must
solve the minimax problem on the left--hand side of eq.\ (\ref{minimax}).
But if eq.\ (\ref{minimax}) holds then there is a saddle point,
and $s^*$ if the first coordinate
of this saddle point, $(s^*,Q^*)$. But then $s^*$ and $Q^*$ must be related
according to the conditions of Theorem 1, as explained above.
When does eq.\ (\ref{minimax}) hold? In general, the well--known sufficient
conditions for
\begin{equation}
\min_{u\in{\cal U}}\max_{v\in{\cal V}}f(u,v)=\max_{v\in{\cal V}}\min_{u\in{\cal U}}f(u,v)
\end{equation}
are that ${\cal U}$ and ${\cal V}$ are convex sets (with ${\cal U}$ being independent
of $v$ and ${\cal V}$ being independent of $u$), and that $f$ is convex in $u$ and concave
in $v$. In our case,
since the function $f(s,Q)=\alpha\bE_Q\ell(X,s)-D(Q\|P)$ is always concave in
$Q$, this sufficient condition would automatically hold whenever
$\ell(x,s)$ is convex in $s$ (for every fixed $x$),
provided that ${\cal S}$ is a space in which
convex combinations can be well defined, and that ${\cal S}$ is a convex set.
\vspace{0.2cm}
\noindent
{\it Maximizing Negative Exponential Moments.}
A similar, but somewhat different, criterion pertaining to exponential
moments, which is reasonable to the same extent,
is the dual problem of $\max_{s\in{\cal S}}\bE\exp
\{-\alpha\ell(X,s)\}$ (again, with $\alpha > 0$). If $\ell(x,s)$
is non-negative for all $x$ and $s$, this has the
advantage that the exponential moment
is finite for all $\alpha > 0$, as opposed to
$\bE\exp\{\alpha\ell(X,s)\}$ which, in many cases, is finite only for
a limited range of $\alpha$. For the same considerations as before, here
we have:
\begin{eqnarray}
\max_s\bE\exp\{-\alpha\ell(X,s)\}&=&
\max_s\exp\{\max_Q[-\alpha\bE_Q\ell(X,s)-D(Q\|P)]\}\nonumber\\
&=&\exp\{-\min_s\min_Q[\alpha\bE_Q\ell(X,s)+D(Q\|P)]\},
\end{eqnarray}
and so the optimality conditions relating $s$ and $Q$ are similar to those
of Theorem 1 (with $\alpha$ replaced by $-\alpha$), except that now we have
a double minimization problem rather than a min--max problem. However, it
should be noted that here
the conditions of Theorem 1 are only
necessary conditions, as for the above equalities to hold,
the pair $(s,Q)$ should {\it globally} minimize
the function $[\alpha\bE_Q\ell(X,s)+D(Q\|P)]$, unlike the earlier
case, where only a saddle point was sought.\footnote{In other
words, it is not enough
now that $s$ and $Q$ are in `equilibrium' in the sense that $s$
is a minimizer for a given $Q$ and vice versa.} On the other hand,
another advantage of this criterion, is that even if one cannot solve
explicitly the equation for the optimum $s$,
then the double minimization naturally suggests
an iterative algorithm: starting from an initial guess $s_0\in{\cal S}$, one
computes $Q_0(x)\propto P(x)\exp\{-\alpha\ell(x,s_0)\}$ (which minimizes
$[\alpha\bE_Q\ell(X,s)+D(Q\|P)]$ over $\{Q\}$), then one finds
$s_1=\mbox{arg}\min_{s\in{\cal S}}\bE_{Q_0}\{\ell(X,s)\}$, and so on. It is
obvious that $\bE\exp\{-\alpha\ell(X,s_i)\}$, $i=0,1,2,\ldots$,
increases (and hence improves)
from iteration to iteration. This is different from the min--max situation
we encountered earlier, where successive improvements are not guaranteed.
\section{A Few Examples}
Theorem 1 tells us that if we are fortunate enough to find a strategy
$s\in{\cal S}$ and a probability distribution $Q$, which are `matched'
to one another (in the sense defined by the above conditions), then we
have solved the problem of minimizing the exponential moment. Sometimes it is
fairly easy to find such a pair $(s,Q)$ by solving an equation. In other
cases, there might be a natural guess for the optimum $s$, which can be
proven optimum by checking the conditions.
In this section, we will see examples of both types.
Some of these examples could have been also solved directly, without using
Theorem 1, but for others, this does not seem to be a trivial task.
In some of the examples, it turns out that the same optimum strategy that
minimizes expected loss, is also optimum in the sense of minimizing all
exponential moments, but this is, of course, not always the case.
\subsection{Example 1: Lossless Data Compression}
We begin with a very simple example.
Let $X$ be a random variable taking on values in a finite alphabet ${\cal X}$,
let $s$ be a probability distribution on ${\cal X}$, i.e., a vector
$\{s(x),~x\in{\cal X}\}$ with $\sum_{x\in{\cal X}}s(x)=1$ and $s(x)\ge 0$ for all
$x\in{\cal X}$, and let $\ell(x,s)\stackrel{\mbox{(d)}}{=} -\ln s(x)$. This example is clearly
motivated by lossless data compression, as $-\ln s(x)$ is the length function
(in nats) pertaining to a uniquely decodable code that is induced by a distribution $s$,
ignoring integer length constraints. In this problem, one readily observes
that the optimum $s$ for minimizing $\bE_Q\{-\ln s(X)\}$ is $s_Q=Q$. Thus, by
eq.\ (\ref{cond}), we seek a distribution $Q$ such that
\begin{equation}
Q(x) \propto P(x)\exp\{-\alpha\ln Q(x)\}=\frac{P(x)}{[Q(x)]^\alpha}
\end{equation}
which means $[Q(x)]^{1+\alpha}\propto P(x)$, or equivalently,
$Q(x)\propto [P(x)]^{1/(1+\alpha)}$. More precisely,
\begin{equation}
s_Q(x)=Q(x)=\frac{[P(x)]^{1/(1+\alpha)}}{\sum_{x'\in{\cal X}}[P(x')]^{1/(1+\alpha)}},
\end{equation}
and the expectation of $-\ln s_Q(X)$
yields the R\'enyi entropy. Note that here
$\ell(x,s)$ is convex in $s$
and so, the minimax condition holds.
While this result is well known and it could have been obtained
even without using Theorem 1, our purpose in this example was to show
how Theorem 1 gives the desired solution even more easily than with the direct
method, by solving
a very simple equation.
\subsection{Example 2: Bayesian Estimation}
Let $(X,Y)$ be random variables, where $Y$ is distributed according to a
given density $P(y)$ and the conditional density of $X$ given $Y$ is given by
\begin{equation}
P(x|y)=\frac{1}{\sqrt{2\pi}}\exp\left\{-\frac{1}{2}(x-\phi(y))^2\right\},
\end{equation}
where $\phi(y)$ is a given function. We are seeking the optimum linear
estimator of $X$ based on the observation $Y$ in the sense of minimizing
the exponential moment of squared error. In other words, we seek a real number
$s$ that minimizes
$\bE\exp\{\alpha(X-sY)^2\}$, where $\alpha\in(0,1/2)$.
Once again, the loss function is convex in $s$.
According to the second condition of Theorem 1, $Q$ should be of the form
\begin{eqnarray}
Q(x,y)&\propto&P(y)\cdot
\frac{1}{\sqrt{2\pi}}\exp\left\{-\frac{1}{2}(x-\phi(y))^2+\alpha(x-sy)^2\right\}\nonumber\\
&=&P(y)\exp\left\{\frac{\alpha}{1-2\alpha}[\phi(y)-sy]^2\right\}\cdot
Q(x|y)\nonumber\\
&\propto&\tilde{P}(y)\cdot
Q(x|y)
\end{eqnarray}
where $Q(x|y)$ is a Gaussian distribution with mean $[\phi(y)-2\alpha
sy]/(1-2\alpha)$ and variance $1/(1-2\alpha)$ and
\begin{equation}
\tilde{P}(y)\propto
P(y)\exp\left\{\frac{\alpha}{1-2\alpha}[\phi(y)-sy]^2\right\}.
\end{equation}
On the other hand, by the first condition of Theorem 1,
$s$ should be the coefficient pertaining to the optimum
linear estimator of $X$ based on $Y$ under $Q$, which is
\begin{equation}
s=\frac{\bE_Q(XY)}{\bE_Q(Y^2)}=\frac{\bE_Q\{Y\cdot\bE_Q(X|Y)\}}{\bE_Q(Y^2)}.
\end{equation}
But since $Q(x|y)$ is Gaussian with mean $[\phi(y)-2\alpha
sy]/(1-2\alpha)$ as said, then this is exactly the inner expectation at the numerator,
and so, we obtain
\begin{equation}
s=\frac{1}{1-2\alpha}\left[\frac{\bE_Q\{Y\cdot\phi(Y)\}}{\bE_Q(Y^2)}-2\alpha
s\right],
\end{equation}
or equivalently,
\begin{equation}
s=\frac{\bE_Q\{Y\phi(Y)\}}{\bE_Q(Y^2)}.
\end{equation}
But since these expectations involve only the random variable $Y$ whose
marginal under $Q$ is $\tilde{P}$, then the
expectations are actually taken under $\tilde{P}$, i.e.,
\begin{equation}
\label{sequation}
s=\frac{\bE_{\tilde{P}}\{Y\phi(Y)\}}{\bE_{\tilde{P}}(Y^2)}.
\end{equation}
Note that this is different from the solution to the ordinary MMSE
problem, where the solution is given by the same expression, but
with $\tilde{P}$ being replaced by $P$. It should be kept in mind that
$\tilde{P}$ depends, in general, on $s$, then so does the right hand side of
the last equation. We have therefore obtained an equation whose solution
$s=s_Q$ is the optimum coefficient in the sense of minimum
$\bE_P\exp\{\alpha(X-sY)^2\}$.
Let us now examine a few simple special cases.
Consider first the case
$\phi(y)=s_0y$, for some real constant $s_0$. In this case, the right
hand--side of eq.\ (\ref{sequation}) is trivially equal to $s_0$, which means
that $s_Q=s_0$. This means that whenever $(X,Y)$ is a Gaussian vector, the
linear MMSE estimator minimizes also
all exponential moments of the squared error (among all linear estimators).
Consider next the case where
\begin{equation}
P(y)=\frac{1}{2}\delta(y-1)+\frac{1}{2}\delta(y+1),
\end{equation}
and denote $\phi_+\stackrel{\mbox{(d)}}{=}\phi(+1)$ and $\phi_-\stackrel{\mbox{(d)}}{=}\phi(-1)$.
Then
\begin{eqnarray}
\tilde{P}(y)&=&\frac{\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_+-s)^2\right\}}
{\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_+-s)^2\right\}+
\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_-+s)^2\right\}}\cdot\delta(y-1)+\nonumber\\
& &\frac{\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_-+s)^2\right\}}
{\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_+-s)^2\right\}+
\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_-+s)^2\right\}}\cdot\delta(y+1).
\end{eqnarray}
Since $\bE_{\tilde{P}}(Y^2)=1$, the equation in $s$ reads
\begin{equation}
s=\frac{\phi_+\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_+-s)^2\right\}-\phi_-
\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_-+s)^2\right\}}
{\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_+-s)^2\right\}+
\exp\left\{\frac{\alpha}{1-2\alpha}(\phi_-+s)^2\right\}}
\end{equation}
For $\phi_+=-\phi_-$, we get $s=\phi_+$, which is expected since this is
actually the linear case discussed above, with $s_0=\phi_+$.
For $\phi_+=\phi_-\stackrel{\mbox{(d)}}{=}\phi$, the equation reads
\begin{equation}
s=-\phi\tanh\left(\frac{2\alpha\phi s}{1-2\alpha}\right)
\end{equation}
and the only solution is $s=0$, which makes sense since $X$ and $Y$ are
independent in this case. For $\alpha\to 0$, the ordinary MMSE linear
estimator is recovered, whose coefficient is $s=(\phi_+-\phi_-)/2$, as
expected.
Returning to the general setting of this example, let us examine what would
happen if we expand the scope and allow a general, non--linear estimator. In this case, we seek a
general function $s(y)$ such that
\begin{equation}
Q(x,y)\propto P(y)\exp\left\{-\frac{1}{2}(x-\phi(y))^2\right\}\cdot
\exp\{\alpha(x-s(y))^2\}.
\end{equation}
If $\alpha\in(0,1/2)$ and we guess $s(y)=\phi(y)$, we obtain
\begin{equation}
Q(x,y)\propto
P(y)\exp\left\{-\left(\frac{1}{2}-\alpha\right)(x-\phi(y))^2\right\}
\end{equation}
for which the conditional mean, $s_Q(y)=\bE_Q(X|Y=y)$, is indeed $\phi(y)$,
and so, the conditions of Theorem 1 are satisfied. It follows then that
for our example, $s(y)=\bE(X|Y=y)=\phi(y)$, minimizes not only the MSE,
but also all exponential moments of the squared error,
$\bE\exp\{\alpha(X-s(Y))^2\}$ for $0< \alpha < 1/2$.
The same idea applies to somewhat more general situations. Let $\rho(t)$ be an
even function, which is monotonically non--decreasing for $t\ge 0$, and
steeply enough so that $\int_{-\infty}^{+\infty}\mbox{d}t e^{-\beta\rho(t)} <
\infty$ for all $\beta > \beta_0$, where $\beta_0 > 0$ is a certain constant.
Suppose
that $P(x,y)\propto P(y)\exp[-\beta\rho(x-\phi(y))]$ for some $\beta >
\beta_0$, and we are interested
in minimizing the exponential moment $\bE\exp\{\alpha\rho(X-s(Y))\}$. Then,
for every $\alpha\in(0,\beta-\beta_0)$, the choice $s(y)=\phi(y)$ leaves $Q(x|y)$
symmetric about $x=\phi(y)$. If $\tau=0$ minimizes
$\int_{-\infty}^{+\infty}\mbox{d}t \rho(t-\tau)e^{-\beta\rho(t)}$ for every
$\beta > \beta_0$ (which is true in many cases), then the estimator $s(y)=\phi(y)$ minimizes
all exponential moments of $\rho(X-s(Y))$. This can be even further
generalized to cases where $P(x|y)\propto \exp[-\beta\rho_1(x-\phi(y))]$ for a
given symmetric function $\rho_1$ that may
be different from the function $\rho$ for
which we wish to minimize the exponential moment.
The above considerations extend also to signal estimation (prediction,
filtering, etc.): Consider two jointly wide--sense stationary Gaussian
processes, $\{(X_n,Y_n)\}$. Given $(...,Y_{-1},Y_0,Y_1,\ldots)$, each $X_t$ is
Gaussian, with conditional mean given by
$\bE\{X_t|...,Y_{-1},Y_0,Y_1,\ldots\}=\sum_{i=-\infty}^\infty h_iY_{t-i}$,
$\{h_i\}$ being the impulse response of the non--causal Wiener filter.
From the same reasons as before,
the exponential moments of the square error are also
minimized by the non--causal Wiener filter.
It is not clear, however, whether the causal Wiener
filter minimizes the exponentiated square error among all causal filters,
unless the non--causal Wiener filter happens to coincide with the causal
one. Optimum linear prediction of Gaussian processes in the ordinary mean
square error sense are also optimum in the mean exponentiated
squared error sense.
\subsection{Example 3: Non--Bayesian Estimation}
Let $X_1,X_2,\ldots,X_n$ be i.i.d.\ Gaussian random variables with mean
$\theta$ and variance $\sigma^2$. It is very well known that among all
unbiased estimators of $\theta$, the one the minimizes the mean square error
(or equivalently, the estimation error variance) is the sample mean
$s(x_1,\ldots,x_n)=\frac{1}{n}\sum_{i=1}^n x_i$. Does the sample mean
estimator also minimize
$\bE\exp\{\alpha[s(X_1,\ldots,X_n)-\theta]^2\}$ among all unbiased estimators
and for all values of $\alpha$ in the allowed range?
Once again, the class ${\cal S}$ of all unbiased estimators is clearly a convex set
and $(s-\theta)^2$ is convex in $s$.
Let us `guess' that the sample mean indeed minimizes also
$\bE\exp\{\alpha[s(X_1,\ldots,X_n)-\theta]^2\}$ and then check whether it satisfies
the conditions of Theorem 1. The corresponding probability measure $Q$, which
will be denoted here by $Q_\theta$, is given by
\begin{eqnarray}
Q_\theta(x_1,\ldots,x_n)&\propto&
\exp\left\{-\frac{1}{2\sigma^2}\sum_{i=1}^n(x_i-\theta)^2+
\alpha\left(\frac{1}{n}\sum_{i=1}^nx_i-\theta\right)^2\right\}\nonumber\\
&=&
\exp\left\{-\frac{1}{2\sigma^2}\sum_{i=1}^n(x_i-\theta)^2+
\alpha\left[\frac{1}{n}\sum_{i=1}^n(x_i-\theta)\right]^2\right\}\nonumber\\
&=&
\exp\left\{-\frac{1}{2\sigma^2}(\bx-\theta\bu)^TW(\bx-\theta\bu)\right\},
\end{eqnarray}
where $\bx\stackrel{\mbox{(d)}}{=}(x_1,\ldots,x_n)^T$, $\bu=(1,1,\ldots,1)^T\in\reals^n$ and
\begin{equation}
W=I-\frac{2\alpha\sigma^2}{n^2}\bu\bu^T,
\end{equation}
$I$ being the $n\times n$ identity
matrix. The maximum likelihood estimator of $\theta$
under $Q_\theta$ is given by
\begin{equation}
s(\bx)=\frac{\bu^TW\bx}{\bu^TW\bu}=\frac{1}{n}\sum_{i=1}^nx_i,
\end{equation}
namely, the
sample mean. It can easily
be shown to achieve the Cram\'er--Rao
lower bound under $Q_\theta$, which is $\sigma^2/(\bu^TW\bu)$.
Thus, the sample mean estimator is an optimum unbiased estimator for
$Q_\theta$ and hence it satisfies the conditions of Theorem 1. The answer to
the question of the previous paragraph is then affirmative. The best
achievable performance, in the exponential moment sense, is given by
\begin{equation}
\bE\exp\left\{\alpha\left(\frac{1}{n}\sum_{i=1}^nX_i-\theta\right)^2\right\}
=\frac{1}{\sqrt{\mbox{det}(W)}}.
\end{equation}
\subsection{Example 4: The Gaussian Joint Source--Channel Coding Problem}
Consider the Gaussian memoryless source
\begin{equation}
P_U(\bu)= (2\pi\sigma_u^2)^{-n/2}
\exp\left\{-\frac{1}{2\sigma_u^2}\sum_{i=1}^nu_i^2\right\}
\end{equation}
and the Gaussian memoryless channel $\by=\bx+\bz$, where the noise
is distributed according to
\begin{equation}
P_Z(\bz)=(2\pi\sigma_z^2)^{-n/2}
\exp\left\{-\frac{1}{2\sigma_z^2}\sum_{i=1}^nz_i^2\right\}.
\end{equation}
In the ordinary joint source--channel coding problem, one seeks an encoder and
decoder that would minimize $D=\frac{1}{n}\sum_{i=1}^n\bE\{(U_i-V_i)^2\}$,
where $\bV=(V_1,\ldots,V_n)$ is the reconstruction at the decoder. It is very well
known that the best achievable distortion, in this case, is given by
\begin{equation}
D=\frac{\sigma_u^2}{1+\Gamma/\sigma_z^2},
\end{equation}
where $\Gamma$ is the maximum power allowed at the transmitter, and
it may be achieved by a transmitter that simply amplifies the source
by a gain factor of $\sqrt{\Gamma/\sigma_u^2}$ and a receiver that implements
linear MMSE estimation of $U_i$ given $Y_i$, on a symbol--by--symbol basis.
What happens if we replace the criterion of expected distortion by the
criterion of the exponential moment on the distortion,
$\bE\exp\{\alpha\sum_i(U_i-V_i)^2\}$? It is natural to wonder whether simple
linear transmitters and receivers, of the kind defined in the previous
paragraph, are still optimum.
The random object $X$, in this example, is
the pair of vectors $(\bU,\bZ)$,
where $\bU$ is the source vector and $\bZ$ is the channel noise vector,
which under $P=P_U\times P_Z$, are independent
Gaussian i.i.d.\ random vectors with zero mean and variances
$\sigma_u^2$ and $\sigma_z^2$, respectively, as said. Our strategy $s$
consists of the choice
of an encoding function $\bx=f(\bu)$ and a decoding function $\bv=g(\by)$.
The class ${\cal S}$ is then the set of all pairs of functions $\{f,g\}$, where
$f$ satisfies the power constraint $\bE_P\{\|f(\bU)\|^2\}\le n\Gamma$.
Condition 2 of Theorem 1 tells us that the modified probability distribution
of $\bu$ and $\bz$ should be of the form
\begin{equation}
\label{qjsc}
Q(\bu,\bz)\propto
P_U(\bu)P_Z(\bz)\exp\left\{\alpha\sum_{i=1}^n[u_i-g_i(f(\bu)+\bz)]^2\right\}
\end{equation}
where $g_i$ is restriction of $g$ to the $i$--th component of $\bv$.
Clearly, if we continue to restrict the encoder $f$ to be linear, with a gain of
$\sqrt{\Gamma/\sigma_u^2}$, which simply exploits the allowed power $\Gamma$, and the
only remaining room for
optimization concerns the decoder $g$, then we are basically
back to the previous example of Bayesian estimation in the Gaussian regime,
and the optimum choice of the decoder
is a linear one, exactly like in the traditional
mean square error case (from the same consideration as in the Bayesian
estimation example). However, once we extend the scope and allow $f$ to be a non--linear encoder,
then the optimum choice of $f$ and $g$ would no longer remain linear like in
the expected distortion case. It is not difficult to see that the conditions of Theorem 1 are no longer
met for any linear functions $f$ and $g$. The key reason is that while
$Q(\bu,\bz)$ of eq.\ (\ref{qjsc}) continues to be Gaussian (though now $U_i$ and $Z_i$ are
correlated) when $f$ and $g$ are linear, the power constraint,
$\bE_P\{\|\bX\|^2\}\le n\Gamma$, when expressed as an expectation w.r.t.\
$Q$, becomes $\bE_Q\{\|f(\bU)\|^2P(\bU)/Q(\bU)\}\le n\Gamma$, but ``power''
function $\|f(\bu)\|^2P(\bu)/Q(\bu)$, with $P$ and $Q$ being Gaussian densities,
is no longer the usual quadratic function of $f(\bu)$ for
which there is a linear encoder and decoder that is optimum.
Another way to see that linear encoders
and decoders are suboptimal, is to consider the
following argument: For a given $n$, the expected exponentiated squared error
is minimized by a joint source--channel coding system, defined over a
super-alphabet of $n$--tuples, with respect to a distortion measure, defined
in terms of a single super--letter, as
\begin{equation}
d(\bu,\bv)=\exp\left\{\alpha\sum_{i=1}^n(u_i-v_i)^2\right\}.
\end{equation}
For such a joint
source--channel coding system to be optimal, the induced channel $P(\bv|\bu)$
must \cite[p.\ 31, eq.\ (2.5.13)]{Berger71} be proportional to
\begin{equation}
P(\bv)\exp\{-\beta d(\bu,\bv)\}=
P(\bv)\exp\left[-\beta\exp\left\{\alpha\sum_i(u_i-v_i)^2)\right\}\right]
\end{equation}
for some $\beta > 0$,
which is the well--known structure of the optimum test channel that attains the
rate--distortion function for the Gaussian source and the above defined
distortion measure.
Had the aforementioned linear system
been optimum, the optimum output distribution $P(\bv)$ would be
Gaussian, and then $P(\bv|\bu)$ would remain proportional to a double
exponential function of
$\sum_i(u_i-v_i)^2$. However, the linear system induces instead a Gaussian channel
from $\bu$ to $\bv$, which is very different, and therefore cannot be optimum.
Of course,
the minimum of $\bE\exp\{\alpha\sum_i(U_i-V_i)^2\}$ can be approached by
separate source- and channel coding, defined on blocks of super--letters
formed by $n$--tuples. The source encoder is an optimum rate--distortion code
for the above defined `single--letter' distortion measure, operating at a
rate close to the channel capacity, and the channel code is
constructed accordingly to support the same rate.
\section{Universal Asymptotically Optimum Strategies}
The optimum strategy for minimizing
$\bE_P\exp\{\alpha\ell(X,s)\}$ depends, in
general, on both $P$ and $\alpha$.
It turns out, however, that this dependence on $P$ and $\alpha$ can sometimes
be relaxed if one gives up the ambition of deriving a strictly optimum
strategy, and resorts to asymptotically optimum strategies.
Consider the case where, instead of one random variable $X$, we have a random
vector $\bX=(X_1,\ldots,X_n)$, governed by an product form probability function
\begin{equation}
P(\bx)=\prod_{i=1}^nP(x_i),
\end{equation}
where each component $x_i$
of the vector $\bx=(x_1,\ldots,x_n)$ takes on values in a finite set
${\cal X}$. If the $\ell(\bx,s)$ grows linearly\footnote{
This happens, for example, when $\ell$ is additive, i.e.,
$\ell(\bx,s)=\sum_{i=1}^nl(x_i,s)$.}
with $n$ for a given empirical
distribution of $\bx$ and a given $s\in{\cal S}$, then it is expected that
the exponential moment $\bE\exp\{\alpha\ell(\bx,s)\}$ would behave, at least
asymptotically, as an exponential function of $n$. In particular, for a given
$s$, the limit
$$\lim_{n\to\infty}\frac{1}{n}\ln\bE_P\exp\{\alpha\ell(\bX,s)\}$$
exists. Let us denote this limit by $E(s,\alpha,P)$. An {\it asymptotically
optimum} strategy is then a strategy $s^*$ for which
\begin{equation}
\label{asymopt}
E(s^*,\alpha,P)\le E(s,\alpha,P)
\end{equation}
for every $s\in{\cal S}$. An asymptotically optimum strategy $s^*$ is called
{\it universal asymptotically optimum} w.r.t.\ a class
${\cal P}$ of probability distributions, if $s^*$ is independent of $\alpha$
and $P$, yet it satisfies eq.\ (\ref{asymopt}) for all $\alpha$
in the allowed range,
every $s\in{\cal S}$, and every $P\in{\cal P}$.
In this section, we take ${\cal P}$ to be the class of all memoryless sources
with a given finite alphabet ${\cal X}$.
We denote by $T_Q$ the type class pertaining to
an empirical distribution $Q$, namely, the set of vectors
$\bx\in{\cal X}^n$ whose empirical distribution is $Q$.
Suppose there exists a strategy $s^*$ and a function
$\lambda:{\cal P}\to\reals$ such that following two conditions hold:
\begin{itemize}
\item[(a)] For every type class $T_Q$ and every $\bx\in T_Q$, $\ell(\bx,s^*)\le
n[\lambda(Q)+o(n)]$,
where $o(n)$ designates a (positive) sequence that tends to zero as
$n\to\infty$.
\item[(b)] For every type class $T_Q$ and every $s\in{\cal S}$,
\begin{equation}
\bigg|T_Q\cap\left\{\bx:~\ell(\bx,s)\ge n[\lambda(Q)-o(n)]\right\}\bigg|\ge e^{-no(n)}|T_Q|.
\end{equation}
\end{itemize}
It is then a straightforward exercise to show, using the method of types, that
$s^*$ is a universal asymptotically optimum strategy w.r.t.\ ${\cal P}$, with
\begin{equation}
E(s^*,\alpha,P)=\max_Q[\alpha\lambda(Q)-D(Q\|P)],
\end{equation}
where condition (a)
supports the direct part and condition (b) supports the converse part.
The interesting point here then is not quite in the last statement, but in the
fact that there are quite a few application examples where these two
conditions hold at the same time.
Before we provide such examples, however, a few words are in order concerning
conditions (a) and (b).
Condition (a) means that there is a choice of
$s^*$, that does
{\it not} depend on $\bx$ or on its type class,\footnote{
As before, $s^*$ is chosen without observing the
data first.} yet the performance of $s^*$, for every $\bx\in T_Q$, ``adapts'' to
the empirical distribution $Q$ of $\bx$ in a way, that according to
condition (b), is ``essentially optimum'' (i.e., cannot be improved
significantly), at least for a
considerable (non--exponential) fraction of the members of $T_Q$.
It is instructive to relate conditions (a) and (b) above to conditions 1 and 2 of
Theorem 1. First, observe that in order to guarantee asymptotic optimality of
$s^*$, condition 2 of Theorem 1 can be somewhat relaxed: For Jensen's
inequality in (\ref{chain}) to remain exponentially tight, it is no longer necessary
to make the random variable $\alpha\ell(\bX,s)+\ln[P(\bX)/Q(\bX)]$ completely
degenerate (i.e., a constant for every realization $\bx$,
as in condition 2 of Theorem 1), but it is enough to keep it
essentially fixed across a considerably large subset of the
dominant type class,
$T_{Q^*}$, i.e., the one whose empirical distribution
$Q^*$ essentially achieves the maximum of $[\alpha\lambda(Q)-D(Q\|P)]$.
Taking $Q^*(\bx)$ to be the memoryless source induced by the dominant $Q^*$,
this is indeed precisely what happens under conditions (a) and (b), which
imply that
\begin{eqnarray}
\alpha\ell(\bx,s^*)+\ln\frac{P(\bx)}{Q^*(\bx)}&\approx&
n\alpha\lambda(Q)+\sum_{i=1}^n\ln\frac{P(x_i)}{Q^*(x_i)}\nonumber\\
&=&n\alpha\lambda(Q)+n\sum_{x\in{\cal X}}Q^*(x)\ln\frac{P(x)}{Q^*(x)}\nonumber\\
&=&n[\alpha\lambda(Q^*)-D(Q^*\|P)],
\end{eqnarray}
for (at least) a non--exponential fraction of the members of $T_{Q^*}$,
namely, a subset of $T_{Q^*}$
that is large enough to maintain the exponential order of the
(dominant) contribution of $T_{Q^*}$ to $\bE\exp\{\alpha\ell(\bx,s^*)\}$.
Loosely speaking, the combination of conditions (a) and
(b) also means then that $s^*$ is essentially
optimum for (this subset of) $T_{Q^*}$, which is a reminiscence of condition 1
of Theorem 1. Moreover, since $s^*$ ``adapts'' to every $T_Q$, in the sense
explained above, then this has the flavor of the max--min
problem discussed in Section 2, where $s$ is allowed to be optimized for each
and every $Q$. Since the minimizing $s$, in the max-min problem, is independent of $P$ and
$\alpha$, this also explains the universality property of such a strategy.
Let us now discuss a few examples. The first example is that of fixed--rate
rate--distortion coding. A vector $\bX$ that emerges from a memoryless source
$P$ is to be encoded by a coding scheme $s$
with respect to a given additive distortion measure, based on a single--letter
distortion measure $d:{\cal X}\times\hat{{\cal X}}\to\reals$, $\hat{{\cal X}}$
being the reconstruction alphabet. Let $D_Q(R)$ denote the
distortion--rate function of a memoryless source $Q$ (with a finite alphabet
${\cal X}$) relative to the single--letter distortion measure $d$ and
let $\ell(\bx,s)$ designate the distortion between the source vector $\bx$ and
its reproduction, using a rate--distortion code $s$. It is not
difficult to see that this example meets conditions (a) and (b) with
$\lambda(Q)=D_Q(R)$: Condition (a) is based on the type covering lemma
\cite[Section 2.4]{CK81},
according to which each type class $T_Q$ can be completely covered by essentially
$e^{nR}$ `spheres' of radius $nD_Q(R)$ (in the sense of $d$), centered at the
reproduction vectors. Thus $s^*$ can be chosen to be a scheme that encodes
$\bx$ in two parts, the first of which is a header that describes the index of
the type class $T_Q$ of $\bx$ (whose description length is
proportional to $\log n$) and the second part encodes
the index of the codeword within $T_Q$, using $nR$ nats. Condition (b)
is met since there is no way to cover $T_Q$ with exponentially less than
$e^{nR}$ spheres within distortion less than $D_Q(R)$.
By the same token, consider the dual problem of
variable--rate coding within a maximum allowed distortion $D$. In this case, every source vector
$\bx$ is encoded by $\ell(\bx,s)$ nats, and
this time, conditions (a) and (b) apply with
the choice $\lambda(Q)=R_Q(D)$, which is the rate--distortion function of $Q$
(the inverse function of $D_Q(R)$). The considerations are similar to those of
the first example.
It is interesting to particularize this example, of variable--rate
coding, to the lossless case, $D=0$ (thus revisiting Example 1), where $R_Q(0)=H_Q$,
the empirical entropy associated with $Q$. In this case, a more refined result can
be obtained, which extends a well known result due to Rissanen
\cite{Rissanen84} in
universal data compression: According to \cite{Rissanen84}, given a length function of
a lossless data compression $\ell(\bx,s)$ ($s$ being the data compression
scheme), and given a parametric class of sources $\{P_\theta\}$, indexed by
a parameter $\theta\in\Theta\subset\reals^k$, a lower bound on $\bE_\theta\ell(\bX,s)$,
that applies to most\footnote{``Most values of $\theta$'' means all values of
$\theta$ with the possible exception of a subset of $\Theta$ whose Lebesgue
measure tends to zero as $n$ tends to infinity.} values
of $\theta$, is given by
\begin{equation}
\bE_\theta \ell(\bX,s)\ge nH_\theta+(1-\epsilon)\frac{k}{2}\log n,
\end{equation}
where $\epsilon > 0$ is arbitrarily small (for large $n$), $H_\theta$ is the entropy
associated with $P_\theta$, and $\bE_\theta\{\cdot\}$ is the expectation under
$P_\theta$. On the other hand, the same expression is achievable, by a number
of universal coding schemes, provided that
the factor $(1-\epsilon)$ in the above expression is replaced by
$(1+\epsilon)$.
Consider now the case where $\{P_\theta,~\theta\in\Theta\}$ is the
class of all memoryless sources over ${\cal X}$, where the parameter vector $\theta$
designates $k=|{\cal X}|-1$ letter probabilities. As for a lower bound, we have
\begin{eqnarray}
\ln\bE_P\exp\{\alpha\ell(\bX,s)\}&\ge&\max_Q\left[\alpha\bE_Q\ell(\bX,s)-nD(Q\|P)\right]\nonumber\\
&\ge&\max_Q\left\{\alpha\left[nH_Q+(1-\epsilon)\frac{k}{2}\ln
n\right]-nD(Q\|P)\right\}\nonumber\\
&=&n\max_Q\left[\alpha H_Q-D(Q\|P)\right]+\alpha(1-\epsilon)\frac{k}{2}\ln n\nonumber\\
&=&n\alpha H_{1/(1+\alpha)}(P)+\alpha(1-\epsilon)\frac{k}{2}\ln n,
\end{eqnarray}
where the second line follows from Rissanen's lower bound (for most sources),
and where $H_u(P)$ is R\'enyi's entropy of order $u$, namely,
\begin{equation}
H_u(P)=\frac{1}{1-u}\ln\left[\sum_{x\in{\cal X}}P(x)^u\right].
\end{equation}
Consider now a two--part code $s^*$, which first encodes the index of the type class
$Q$ and then the index of $\bx$ within the type class. The corresponding
length function is given by
\begin{equation}
\ell(\bx,s^*)=\ln|T_Q|+k\ln n\approx n\hat{H}(\bx)+\frac{k}{2}\ln n,
\end{equation}
where $\hat{H}(\bx)$ is the empirical entropy pertaining to $\bx$, and where
the approximate inequality is easily obtained by the Sterling approximation.
Then,
\begin{eqnarray}
\ln\bE_P\exp\{\alpha\ell(\bX,s)\}&=&\ln\bE\exp\{\alpha
n\hat{H}(\bX)\}+\alpha\frac{k}{2}\ln n\nonumber\\
&=&\ln\bE_P\exp\{\alpha \min_Q[-\ln Q(\bX)]\}+\alpha\frac{k}{2}\ln
n\nonumber\\
&\le&\min_Q\ln\bE_P\exp\{-\alpha\ln Q(\bX)\}+\alpha\frac{k}{2}\ln n\nonumber\\
&=&n\alpha H_{1/(1+\alpha)}(P)+\alpha\frac{k}{2}\ln n,
\end{eqnarray}
and then it essentially achieves the lower bound.
Rissanen's result is now a special case
of this, corresponding to $\alpha\to 0$.
Our last example corresponds to a secrecy system. A sequence $\bx$ is to be
communicated to a legitimate decoder which shares with the transmitter a
random key $\bz$ of $nR$ purely random bits. The encoder transmits an
encrypted message $\by=\phi(\bx,\bz)$, which is an invertible function of
$\bx$ given $\bz$, and hence decipherable by the legitimate decoder. An
eavesdropper, which has no access to the key $\bz$, submits a sequence of guesses
concerning $\bx$ until it receives an indication that the last guess was
correct (e.g., a correct guess of a password admits the eavesdropper into a secret
system). For the best possible encryption function $\phi$, what would be the
optimum guessing strategy $s^*$ that the eavesdropper may apply
in order to minimize the $\alpha$--th moment of the number of
guesses $G(\bX,s)$, i.e., $\bE\{G^\alpha(\bX,s)\}$? In this case,
$\ell(\bx,s)=\ln G(\bx,s)$. As is shown in \cite{MA99}, there exists a
guessing strategy $s^*$, which for every $\bx\in T_Q$, gives
$\ell(\bx,s^*)\approx n\min\{H_Q,R\}$, a quantity that essentially cannot be
improved upon by any other guessing strategy, for most members of $T_Q$. In other
words, conditions (a) and (b) apply with $\lambda(Q)=\min\{H_Q,R\}$.
\section{Phase Transitions}
Another interesting aspect of the asymptotic behavior of the exponential
moment is the possible appearance of phase transitions, i.e.,
irregularities in the exponent function
$E(s,\alpha,P)$ even in some very simple and `innocent' models.
By irregularities, we mean a non--smooth behavior, namely,
discontinuities in the derivatives of $E(s,\alpha,P)$ with respect to $\alpha$
and/or the parameters of the source $P$.
One example that exhibits phase
transitions is that of the secrecy system,
mentioned in the last paragraph of the previous section. As is shown in
\cite{MA99}, the optimum exponent $E(s^*,\alpha,p)$ for this case
consists of two phase transitions as a function of $R$ (namely, three different
phases). In particular,
\begin{equation}
E(s^*,\alpha,P)=\left\{ \begin{array}{ll}
\alpha R & R < H(P) \\
(\alpha-\theta_R)R+\theta_R H_{1/(1+\theta_R)}(P) & H(P)\le R\le H(P_\alpha) \\
\alpha H_{1/(1+\alpha)}(P) & R > H(P_\alpha)
\end{array} \right.
\end{equation}
where $P_\alpha$ is the distribution defined by
\begin{equation}
P_\alpha(x)\stackrel{\mbox{(d)}}{=}\frac{P^{1/(1+\alpha)}(x)}{\sum_{x'\in{\cal X}} P^{1/(1+\alpha)}(x')},
\end{equation}
$H(Q)$ is the Shannon entropy associated with a distribution $Q$,
$H_u(Q)$ is the R\'enyi entropy of order $u$ as defined before, and
$\theta_R$ is the unique solution of
the equation $R=H(P_\theta)$ for $R$ in the range
$H(P) \le R \le H(P_\alpha)$.
But this example may not really be extremely
surprising due to the non--smoothness of the function
$\lambda(Q)=\min\{H_Q,R\}$.
It may be somewhat less expected, however, to witness phase
transitions also in some very simple and `innocent' looking models.
One way to understand the
phase transitions in these cases,
comes from the statistical--mechanical perspective. It turns
out that in some cases, the expression of the exponential moment is analogous
to that of a partition function of a certain many--particle physical
system with interactions, which may exhibit phase transitions and these phase
transitions correspond the above--mentioned irregularities.
We now demonstrate a very simple model, which has phase transitions.
Consider the case where $\bX$ is a binary vector whose components take
on values in ${\cal X}=\{-1,+1\}$, and which is governed by a binary memoryless
source $P_\mu$ with probabilities
$\mbox{Pr}\{X_i=+1\}=1-\mbox{Pr}\{X_i=-1\}=(1+\mu)/2$ ($\mu$ designating the
expected `magnetization' of each binary spin $X_i$, to make the physical
analogy apparent). The probability of $\bx$ under $P_\mu$ is thus easily
shown to be given by
\begin{equation}
P_\mu(\bx)=\left(\frac{1+\mu}{2}\right)^{(n+\sum_ix_i)/2}\cdot
\left(\frac{1-\mu}{2}\right)^{(n-\sum_ix_i)/2}
=\left(\frac{1-\mu^2}{4}\right)^{n/2}\cdot\left(\frac{1+\mu}{1-\mu}\right)^{\sum_ix_i/2}.
\end{equation}
Consider the estimation of the parameter $\mu$ by the ML estimator
\begin{equation}
\hat{\mu}=\frac{1}{n}\sum_{i=1}^nx_i.
\end{equation}
How does the exponential moment of $\bE_\mu\exp\{\alpha n(\hat{\mu}-\mu)^2\}$
behave like? A straightforward derivation yields
\begin{eqnarray}
\bE_\mu\exp\{\alpha n(\hat{\mu}-\mu)^2\}&=&
\left(\frac{1-\mu^2}{4}\right)^{n/2}e^{n\alpha\mu^2}
\sum_{\bx}\left(\frac{1+\mu}{1-\mu}\right)^{\sum_ix_i/2}\exp\left\{\frac{\alpha}{n}\left(\sum_ix_i\right)^2
-2\alpha\mu\sum_ix_i\right\}\nonumber\\
&=&\left(\frac{1-\mu^2}{4}\right)^{n/2} e^{n\alpha\mu^2}
\sum_{\bx}\exp\left\{\left(\frac{1}{2}\ln\frac{1+\mu}{1-\mu}-2\alpha\mu\right)\sum_ix_i+
\frac{\alpha}{n}\left(\sum_ix_i\right)^2\right\}.\nonumber
\end{eqnarray}
The last summation over $\{\bx\}$ is exactly the partition function pertaining
to the {\it Curie--Weiss model} of spin arrays in statistical mechanics (see, e.g.,
\cite[Subsection 2.5.2]{MM09}), where
the magnetic field is given by
\begin{equation}
B=\frac{1}{2}\ln\frac{1+\mu}{1-\mu}-2\alpha\mu
\end{equation}
and the coupling coefficient for every pair of spins is $J=2\alpha$.
It is well known that this model exhibits phase transitions pertaining to
spontaneous magnetization below a certain critical temperature. In particular,
using the method of types \cite{CK81}, this partition function can be asymptotically evaluated
as being of the exponential order of
$$\exp\left\{n\cdot\max_{|m|\le
1}\left[h_2\left(\frac{1+m}{2}\right)+Bm+\frac{J}{2}\cdot m^2\right]\right\},$$
where $h_2(\cdot)$ is the binary entropy function, which stands for the
exponential order of the number of configurations $\{\bx\}$ with a given
value of $m=\frac{1}{n}\sum_ix_i$. This expression
is clearly dominated by a
value of $m$ (the dominant magnetization $m^*$)
which maximizes the expression in the square brackets, i.e., it solves the equation
\begin{equation}
m=\tanh(Jm+B),
\end{equation}
or in our variables,
\begin{equation}
m=\tanh\left(2\alpha m+\frac{1}{2}\ln\frac{1+\mu}{1-\mu}-2\alpha\mu\right).
\end{equation}
For $\alpha < 1/2$, there is only one solution and there is no spontaneous
magnetization (paramagnetic phase). For $\alpha > 1/2$, however, there are
three solutions, and only one of them dominates the partition function,
depending on the sign of $B$, or equivalently, on whether $\alpha >
\alpha_0(\mu)\stackrel{\mbox{(d)}}{=}\frac{1}{4\mu}\ln\frac{1+\mu}{1-\mu}$ or $\alpha <
\alpha_0(\mu)$ and according to the sign of $\mu$. Accordingly, there are five
different phases in the plane spanned by $\alpha$ and $\mu$. The paramagnetic
phase $\alpha < 1/2$, the phases $\{\mu > 0,~1/2 < \alpha < \alpha_0(\mu)\}$
and $\{\mu < 0,~ \alpha > \alpha_0(\mu)\}$, where the dominant magnetization
$m$ is positive, and the two complementary phases,
$\{\mu < 0,~1/2 < \alpha < \alpha_0(\mu)\}$ and
and $\{\mu > 0,~ \alpha > \alpha_0(\mu)\}$, where the dominant magnetization
is negative. Thus, there is a multi-critical point
where the boundaries of all five phases meet, which the point
$(\mu,\alpha)=(0,1/2)$. The phase diagram is depicted in Fig.\ 1.
\begin{figure}[ht]
\hspace*{3cm}\input{phd.pstex_t}
\caption{\small Phase diagram in the plane of $(\mu,\alpha)$.}
\label{mm1}
\end{figure}
Yet another example of phase transitions is that of fixed--rate lossy data
compression, discussed in the previous section.
To demonstrate this explicitly, consider the binary symmetric
source (BSS) and the Hamming distortion measure $d$, and consider a random
selection of a rate--$R$ code
by $ne^{nR}$ independent fair coin tosses, one for each of the $n$ components
of every one of the $e^{nR}$ codewords. It was shown in \cite{Merhav09} that
the asymptotic exponent of the negative exponential moment,
$\bE\exp\{-\alpha\sum_id(U_i,V_i)\}$ (where
the expectation is w.r.t.\ both the source and the random code selection),
is given by the following expression, which obviously exhibits a (second
order) phase transition:
\begin{equation}
\lim_{n\to\infty}\frac{1}{n}\ln
\bE\exp\left\{-\alpha\sum_id(U_i,V_i)\right\}
=\left\{\begin{array}{ll}
-\alpha\delta (R) & \alpha\le \alpha(R)\\
-\alpha+\ln(1+e^\alpha)+R-\ln 2 & \alpha > \alpha(R)
\end{array}
\right.
\end{equation}
where $\delta(R)$ is the distortion--rate function of the BSS w.r.t.\
the Hamming distortion measure and
\begin{equation}
\alpha(R)=\ln\frac{1-\delta(R)}{\delta(R)}.
\end{equation}
The analysis in \cite{Merhav09} is based on the random energy model (REM),
\cite{Derrida80},\cite{Derrida80b},\cite{Derrida81},
a well-known statistical--mechanical model of spin glasses with
strong disorder, which is known to exhibit phase transitions. Moreover, it is
shown in \cite{Merhav09} that ensembles of codes that have an hierarchical
structure may have more than one phase transition.
\section{Lower Bounds on Exponential Moments}
As explained in the Introduction,
even in the ordinary setting, of the quest for minimizing
$\bE\{\ell(X,s)\}$, optimum strategies may not always be known, and then
useful lower bounds are very important.
This is definitely the case when exponential moments are
considered, because the exponential moment criterion is even harder to handle. To obtain non--trivial
bounds on exponential moments, we propose to harness lower bounds the expectation of $\ell(X,s)$,
possibly using a change of measure, in the spirit of the proof of Theorem 1
and the previous example of a lower bound on universal lossless data
compression. We next demonstrate this idea in the context of a lower bound on the expected
exponentiated squared error of an unbiased estimator, on the basis of the Cram\'er--Rao bound (CRB).
The basic idea, however, is applicable more generally, e.g., by relying on
other well--known Bayesian/non--Bayesian bounds on the mean-square error (e.g., the Weiss--Weinstein
bound for Bayesian estimation \cite{WW85}), as well as in
bounds on signal estimation (filtering, prediction, etc.),
and in other problem areas as well. Further investigation in the line may be of
considerable interest.
Consider a parametric family of probability distributions
$\{P_\theta,~\theta\in\Theta\}$, $\Theta\subseteq\reals$
being the parameter set, and suppose that we are interested in
a lower bound on $\bE_\theta\exp\{\alpha(\hat{\theta}-\theta)^2\}$, for
any unbiased estimator of $\theta$, where as before, $\bE_\theta$ denotes
expectation w.r.t.\ $P_\theta$. Consider the following chain of inequalities,
which holds for any $\theta'\in\Theta$:
\begin{eqnarray}
\bE_\theta\exp\{\alpha(\hat{\theta}-\theta)^2\}&=&
\bE_{\theta'}\exp\left\{\alpha(\hat{\theta}-\theta)^2+\ln\frac{P_\theta(X)}{P_{\theta'}(X)}\right\}\nonumber\\
&\ge&\exp\left\{\alpha\bE_{\theta'}(\hat{\theta}-\theta)^2-D(P_{\theta'}\|P_\theta)\right\}\nonumber\\
&=&\exp\left\{\alpha\bE_{\theta'}(\hat{\theta}-\theta')^2+
\alpha(\theta-\theta')^2-D(P_{\theta'}\|P_\theta)\right\}\nonumber\\
&\ge&\exp\left\{\alpha\mbox{CRB}(\theta')+
\alpha(\theta-\theta')^2-D(P_{\theta'}\|P_\theta)\right\},
\end{eqnarray}
where $\mbox{CRB}(\theta)$ is the Cram\'er--Rao bound for unbiased estimators, computed at $\theta$ (i.e.,
$\mbox{CRB}(\theta)=1/I(\theta)$, where $I(\theta)$ is the Fisher information). Since
this lower bound applies for every $\theta'\in\Theta$, one can take its
supremum over $\theta'\in\Theta$ and obtain
\begin{equation}
\ln\bE_\theta\exp\{\alpha(\hat{\theta}-\theta)^2\}\ge
\sup_{\theta'\in\Theta}\left[\alpha\mbox{CRB}(\theta')+\alpha(\theta'-\theta)^2-D(P_{\theta'}\|P_\theta)\right].
\end{equation}
More generally if $\theta=(\theta_1,\ldots,\theta_k)^T$
is a parameter vector (thus
$\theta\in\Theta\subseteq\reals^k$) and $\alpha\in\reals^k$ is an arbitrary
deterministic (column) vector, then
\begin{equation}
\ln\bE_\theta\exp\{\alpha^T(\hat{\theta}-\theta)(\hat{\theta}-\theta)^T\alpha\}\ge
\sup_{\theta'\in\Theta}\left[\alpha^TI^{-1}(\theta')\alpha+[\alpha^T(\theta'-\theta)]^2-
D(P_{\theta'}\|P_\theta)\right],
\end{equation}
where here $I(\theta)$ is the Fisher information matrix and $I^{-1}(\theta)$
is its inverse.
It would be interesting to further investigate bounds of this type, in
parameter estimation in particular, and in other problem areas in general,
and to examine when these bounds may be tight and useful.
\section*{Acknowledgment}
Interesting discussions with Rami Atar are acknowledged with thanks.
|
\section{Introduction}
\label{sec:introdetrip}
\subsection{Compressed Sensing}
\label{subsec:compressedsensing}
Many signal processing
applications focus on identifying and estimating
a few significant coefficients from a high dimension vector.
The wisdom behind this approach is the ubiquitous compressibility of
signals:
most of the information
contained in a signal often resides in just a few large coefficients.
Traditional sensing, compression and processing systems first acquire the entire data,
apply a transformation to the data,
and then discard most of the coefficients;
we retain only a small number of
significant coefficients.
Clearly, it is wasteful to sense and compute on
all of the coefficients
when most coefficients will be discarded at a later stage.
This naturally begs the question: can we sense compressible signals in a
compressible way? In other words, can we sense only that
portion of the signal that will not be thrown away?
The ground-breaking work of compressed sensing (CS)
pioneered by
Cand\'{e}s et al.~\cite{CandesRUP} and
Donoho~\cite{DonohoCS} answers this question in the affirmative.
Cand\'{e}s et al.~\cite{CandesRUP} and
Donoho~\cite{DonohoCS} have demonstrated that the information contained in the few
significant coefficients can be captured (encoded)
by a small number of \emph{random linear projections}.
The original signal can then be reconstructed (decoded)
from these random projections using
an appropriate decoding scheme.
Consider a discrete-time signal $x\in{\mathbb R}^n$
that has only $k \ll n$ non-zero coefficients.
CS posits that
it is unnecessary to measure all the $n$
values of $x$;
rather,
we can recover $x$ from a small number of projections onto
an {\em incoherent} basis \cite{CandesRUP,DonohoCS}.
To measure (encode) $x$, we
compute the measurement vector $y \in {\mathbb R}^m$
containing $m$ linear projections of $x$ via
the matrix-vector multiplication $y=\Phi x$, where $\Phi\in{\mathbb R}^{m \times n}$
is the CS matrix.
The CS theory asserts that we can reconstruct (decode) $x$
given $y$ and $\Phi$ using $m \ll n$
measurements, provided certain requirements on $\Phi$ are satisfied.
It can be shown that if the CS matrix $\Phi$ is constructed by filling its $m \times n$ entries randomly from an i.i.d. Gaussian distribution, then with probability one, $m=k+1$ measurements are sufficient to encode $x$. In particular, with probability one, $x$ can be reconstructed exactly from $y \in {\mathbb R}^m$, $m \ge k+1$,
using $\ell_0$ minimization \cite{DCS}:
\begin{equation}
\min_{x \in {\mathbb R}^{n}} \|x\|_0 \mbox { subject to } \Phi x = y,
\label{eq:l0minimization}
\end{equation}
where $\|x\|_0 = \{\#x_i : x_i \ne 0 \}$.
However, signal recovery algorithms using as few as $m=k+1$ measurements require
a search in each of the ${n \choose k}$ subspaces that could contain the significant signal coefficients.
Consequently the complexity of the algorithm to recover $x$ using (\ref{eq:l0minimization}) is NP complete \cite{CandesECLP}.
Fortunately, at the expense of
acquiring slightly more measurements, we can recover $x$ from $y$ thru a
convex relaxation of (\ref{eq:l0minimization}); the complexity of the resulting recovery algorithm can be made polynomial. With $m \approx k \log(n/k)$ measurements, the solution to the $\ell_1$ minimization (which can be solved with cubic complexity) given by
\begin{equation}
\min_{x \in {\mathbb R}^{n}} \|x\|_1 \mbox { subject to } \Phi x = y,
\label{eq:l1minimization}
\end{equation}
coincides with the solution of (\ref{eq:l0minimization}) for $\Phi$ constructed from i.i.d. Gaussian distribution
\cite{CandesRUP,DonohoCS}. The ability to recover sparse signals easily (in polynomial time) from the small number of CS measurements is one of the main reasons why CS has enjoyed tremendous attention in the research community over the last few years \cite{CSweb}.
\subsection{What is a Good Compressed Sensing Matrix?}
The CS matrix $\Phi$ plays a vital role in both data acquisition and the subsequent recovery of sparse signals.
Not only do the properties of $\Phi$ dictate how much information we capture about the signal $x$, but they also
determines the ease of reconstructing $x$ from the measurements $y$.
In this paper, we select Restricted Isometry as proposed by
Cand\'{e}s and Tao \cite{CandesDLP} as the metric to determine whether a given $\Phi$ is a good candidate
for CS data acquisition.
The reason we
choose this metric is because several key results in CS depend on the Restricted Isometry properties of $\Phi$
\cite{CandesDLP,CandesRUP,jlpaper,gurevich,CalderbankDetRIP,blanchard2010compressed,blanchard2009decay,bah2010improved,blanchard2010support,chartrand2008restricted,garg2009gradient,davenport2010analysis,haupt2007generalized,Candes2008}.
\subsection{Restricted Isometry Property}
\begin{DEFI}
For each integer $k=1,2,...$, define the {\em Restricted Isometry constant} $\delta_k$ of a matrix $\Phi \in {\mathbb R}^{m \times n}$ as the smallest number such that
\begin{equation}
(1-\delta_k) \|x\|_{\ell_2}^{2} \le \| \Phi x \|_{\ell_2}^{2} \le (1+\delta_k) \|x\|_{\ell_2}^{2}
\label{eq:riconstant}
\end{equation}
holds for all non-zero vectors $x \in {\mathbb R}^n$ that satisfy $0 < \|x\|_0 \le k$.
\end{DEFI}
Note that a ``good'' CS matrix has a small Restricted Isometry constant $\delta_k$.
As noted before, a number of key results in CS involve the Restricted Isometry
constant.
We highlight two results,
both taken from \cite{Candes2008}.
Theorem \ref{th:l0l1equal} shows that if the Restricted Isometry constant $\delta_{2k}$ of order $2k$ is
sufficiently small, then CS recovery is guaranteed to be tractable.
Theorem \ref{th:noisycs} shows that the same condition on $\delta_{2k}$ guarantees robustness in CS recovery when the measurements are
corrupted with bounded additive noise.
For Theorems \ref{th:l0l1equal} and \ref{th:noisycs}, we assume $x$ is {\em any} signal in ${\mathbb R}^n$ and not necessarily $k$-sparse. Let $x_k \in {\mathbb R}^n$ be the
best $k$-term approximation of $x$, ie, $x_k$ is obtained by taking $x$ and setting all but the $k$ largest magnitude entries to zero. Note that if $x$ is $k$-sparse, then $x=x_k$.
\begin{THEO} {\em \cite[Theorem 1.1, page 2]{Candes2008}}
If $\delta_{2k}$ satisfies $\delta_{2k} < \sqrt{2}-1$, then the solution $x^{*}$ to the $\ell_1$ minimization (\ref{eq:l1minimization}) obeys
\begin{equation}
\| x^{*} - x\|_{\ell_1} \le C_0 \| x - x_{k}\|_{\ell_1},
\end{equation}
and
\begin{equation}
\| x^{*} - x\|_{\ell_2} \le C_0 k^{-1/2} \| x - x_{k}\|_{\ell_1}
\end{equation}
for some constant $C_0$. In particular, if $x$ is $k$-sparse, then the recovery is exact.
\label{th:l0l1equal}
\end{THEO}
Theorem~\ref{th:l0l1equal} asserts that the solution of the $\ell_1$ minimization is exact when $x$ is $k$-sparse and when $\delta_{2k}$ is sufficiently small.
Rather than solving the intractable $\ell_0$ minimization of (\ref{eq:l0minimization}) directly, we obtain the same solution using the tractable $\ell_1$ minimization as given by (\ref{eq:l1minimization}).
Note that with only $m=k+1$ measurements, the CS matrix cannot satisfy the condition on $\delta_{2k}$. However, with more measurements and with specific construction of
CS matrices
(which can be either deterministic or stochastic), it can be shown that the condition in Theorem \ref{th:l0l1equal} can be satisfied surely or with high probability~\cite{DonohoCS,CandesRUP,jlpaper,CalderbankDetRIP}.
The same upper-bound on $\delta_{2k}$ serves as a sufficient condition for the recovery to be robust in the presence of
bounded additive noise. Consider noisy measurements
\begin{equation}
y = \Phi x + z,
\end{equation}
where $z$ is an unknown noise vector that is bounded $\|z\|_{\ell_2} \le \epsilon$.
Consider the convex optimization problem
\begin{equation}
\min_{x \in {\mathbb R}^{n}} \|x\|_{\ell_1} \mbox { subject to } \|y-\Phi x\|_{\ell_2} \le \epsilon.
\label{eq:l1noiseminimization}
\end{equation}
The following Theorem shows that sufficiently small Restricted Isometry constant $\delta_{2k}$ guarantees robust CS recovery of $k$-sparse
signals in the presence of
bounded measurement noise. Specifically, if the noise is bounded, then the error in signal recovery is also bounded.
\begin{THEO}{\em \cite[Theorem 1.2, page 3]{Candes2008}}
If $\delta_{2k} < \sqrt{2} - 1$ and $\|z\|_{\ell_2} \le \epsilon$, then the solution $x^*$ to the optimization (\ref{eq:l1noiseminimization}) obeys
\begin{equation}
\| x^{*} - x\|_{\ell_2} \le C_0 k^{-1/2} \| x - x_{k}\|_{\ell_1} + C_1 \epsilon
\end{equation}
with the same constant $C_0$ as in Theorem~\ref{th:l0l1equal} and some fixed $C_1$. In particular, if $x$ is $k$-sparse, then
$\| x^{*} - x\|_{\ell_2} \le C_1 \epsilon$.
\label{th:noisycs}
\end{THEO}
The central role played by the Restricted Isometry in CS begs the question: what is the best (smallest) Restricted Isometry constant we can hope to attain for a
CS matrix in ${\mathbb R}^{m \times n}$
for a given triplet $n$, $m$ and $k$?
This key question is what we investigate in this paper.
\subsection{Notations and Definitions}
We define {\em CS problem size} as the triad of numbers
$\{n,m,k\}$. For a fixed problem size, the CS matrix $\Phi$ is chosen from ${\mathbb R}^{m \times n}$.
As we shall see shortly, the Restricted Isometry of $\Phi$
depends on the singular values of the $m \times k$-sized submatrices of $\Phi$ (which are ${n \choose k}$ in number.)
While we focus mainly on real valued $\Phi$,
several of the results we derive hold for complex valued $\Phi$ as well; we will explicitly mention if the results are applicable in the complex domain.
We exploit the fact that the measurements $y$ depend only on
$k$ out of the $n$ columns of $\Phi$ when $x$ is $k$-sparse; the $k$ columns are
the ones that correspond to the indices of the non-zero entries in $x$.
As a consequence, the Restricted Isometry constant depends directly on the properties
of the $m \times k$ sized submatrices of $\Phi$.
This observation motivates us to consider the singular values
of submatrices $\Phi_p \in {\mathbb R}^{m \times k}$ formed by
selecting only $k$ of the $n$ columns of $\Phi$.
The index $p \in [1,2,...,{n \choose k}]$ uniquely identifies the set of
columns chosen from the complete matrix $\Phi$.
Let the singular values of the matrix $\Phi$ be
$S_1$, $S_2$,..., $S_m$, where
$S_1 \ge S_2 \ge...\ge S_m \ge 0$.\footnote{We
restrict our attention only to the largest $m$ singular values of $\Phi$
because the remaining $n-m$ singular values are zero.}
Furthermore, let the singular values of the submatrix $\Phi_p$ be
$s_{p,1}$, $s_{p,2}$,..., $s_{p,k}$, where
$s_{p,1} \ge s_{p,2} \ge...\ge s_{p,k} \ge 0$.
We define $\rho_{\max}(\Phi,k) \ge 0$ and $\rho_{\min}(\Phi,k) \ge 0$, respectively, as the maximum and the minimum
of the singular values of every submatrix of $\Phi$ of size $m \times k$, that is,
\begin{equation}
\rho_{\max}(\Phi,k)= (\max_p \{s_{p,1}\})^2
\mbox{ ~~~~~ and ~~~~~}
\rho_{\min}(\Phi,k)= (\min_p \{s_{p,k}\})^2.
\end{equation}
Clearly, for a $k$-sparse signal $x$, we have that
\begin{equation}
\rho_{\min}(\Phi,k) \|x\|_{\ell_2}^{2} \le \| \Phi x \|_{\ell_2}^{2} \le \rho_{\max}(\Phi,k) \|x\|_{\ell_2}^{2},
\label{eq:tightrhobound}
\end{equation}
where both inequalities in the above equation are tight.
We define the {\em Restricted Isometry Property Ratio} (RIP ratio) as
\begin{equation}
R(\Phi,k) = \frac{\rho_{\max}(\Phi,k)}{ \rho_{\min}(\Phi,k)}.
\label{eq:ripratio}
\end{equation}
This ratio plays the role of the square of the condition number of $\Phi$ when it is
applied exclusively to the domain of $k$-sparse signals.
We denote the differential operator $\frac{d}{dx}$ by $D$, i.e., $D[f(x)]=\frac{d}{dx} f(x)$,
$D^2[f(x)]=\frac{d^2}{dx^2} f(x)$ and so on. We use the same variable name $x$ when we consider polynomials such as the
characteristic equations of $\Phi$ and its submatrices. Any ambiguity with reference to our $k$-sparse signal $x$ is resolved
from the context in which it is presented.
\subsection{Contributions}
Existing results on Restricted Isometry in the CS literature provide achievability results on specific constructions of $\Phi$ (such as random i.i.d. Gaussian) \cite{CandesRUP,DonohoCS,jlpaper,CalderbankDetRIP}.
Most of the results are stochastic, along the lines of: ``constructing a matrix according to a prescribed method
(such as populating the matrix with i.i.d. Gaussian or i.i.d. Bernoulli entries)
yields a CS matrix that satisfies RIP with a given $\delta_k$
with high probability \cite{jlpaper}''. There have also been recent results on deterministic achievability, where a deterministic construction of $\Phi$ is
proved to satisfy a {\em statistical} RIP (more on statistical RIP in Section~\ref{sec:stochrip})
for some constant $\delta_k$ \cite{CalderbankDetRIP}.
Optimal CS reconstruction is intimately related to Gelfand and Kolmogrov widths of $\ell_p$ balls, an area
that was extensively studied in the late 1970's and early 1980's by Kashin, Gluskin, and Garnaev
\cite{Gluskin82,Gluskin84_1,Gluskin84_2,Kashin77_1,Kashin77_2}.
Connections between the two fields have been recognized by Donoho~\cite{DonohoCS} and
others~\cite{remcs,Kashin2007,jlpaper,Foucart2010} and have led to several deterministic bounds on Restricted Isometry. However, these prior results offer little insight into the structure of CS matrices that are optimal in the Restricted Isometry sense. In this work, we fill this gap by revealing the
intricate relationships between optimal CS matrices and several well known results in coding theory and
frame theory. As an important consequence of our work, we stumble upon a result that extends the
well-known Welch bound~\cite{WelchEquiangular} to higher orders.
In this paper, we derive two deterministic converse bounds for the RIP ratio based respectively on the structural
properties of the CS matrix and packing of subspaces in ${\mathbb R}^{m \times n}$.
Additionally, we also derive a deterministic achievable bound for
the RIP ratio, which says that we can {\em certainly} find a $\Phi \in {\mathbb R}^{m \times n}$ that has a better RIP
ratio than our new achievable bound.
The first bound we derive is a deterministic converse bound on the RIP ratio, called the {\em structural bound}. It is a function of the problem size $\{n,m,k\}$.
The key insight that we use to derive it
is that the singular values $\{s_{p,i}\}$ of $\Phi_p$ (submatrices of $\Phi$) are
severely constrained by the singular values $\{S_j\}$ of the complete matrix $\Phi$.
For example, the interlacing inequality \cite{RCThompson9} requires that
\begin{equation*}
s_{p,i} \le S_i \mbox { for all } i=1,2,...,k,
\end{equation*}
and therefore the $\{s_{p,i}\}$'s cannot take on arbitrary values.
We explore the intricate relationships between the two sets $\{s_{p,i}\}$ and $\{S_i\}$ of singular values
to derive the structural bound on the RIP ratio. The structural bound places a limit on the best (smallest)
RIP ratio that is attainable for matrices in
${\mathbb R}^{m \times n}$. Theorem \ref{theo:main1} in Section \ref{sec:mainresults} is the main result on this bound. The Theorem holds not only for
real valued $\Phi$, but also for complex valued $\Phi$. We also show how the structural bound is related to equi-angular tight frames (ETF)~\cite{TroppEquiangular}, Welch bound~\cite{WelchEquiangular}, and the
the Generalized Pythagorean Theorem~\cite{GPT1,GPT2}.
The second bound is also a deterministic converse bound on the RIP, called the {\em packing bound}. It is applicable to real-valued $\Phi$.\footnote{Although
we derive the packing bound for real valued $\Phi$ in this paper, we can use similar arguments to
easily derive another packing bound that applies for complex valued $\Phi$.} In this paper, we derive the packing bound for the limited case of $k=2$; however,
we plan to extend the results for $k>2$ in a future paper.
The need for the packing bound is motivated by the fact that the structural bound is loose for large values of $n$ (in comparison to $m$ and $k$); the packing bound
offers a tighter bound in the regime of large $n$. The packing bound is derived from an entirely different perspective: we consider the $n$ columns of
$\Phi$ as vectors in ${\mathbb R}^m$ and show that minimizing the RIP ratio is equivalent to finding
$n$ vectors in ${\mathbb R}^m$ where the angular separation of every pair of vectors is as large as possible. In other words, we show that
deriving the optimal RIP ratio for $k=2$ is equivalent to optimal packing in Grassmannian spaces. The main result of the packing bound is presented in Theorem~\ref{theo:packing}
in Section \ref{sec:mainresults}.
The two converse bounds are applicable to every matrix in ${\mathbb R}^{m \times n}$. Therefore, the results provide bounds on the best (smallest) possible RIP ratio that any CS matrix can achieve for a given problem size. We also discuss which of the two converse bounds presented is likely to be the tighter bound, based on the particular selection of $\{k,m,n\}$.
The third bound we derive is a deterministic achievable bound, called the {\em covering bound}, which also exploits the equivalence of CS matrices to
packing in a grasmannian space. We use a Theorem derived independently by Chabauty, Shannon, and Wyner
\cite{Chabauty53,Shannon59,Wyner65} that uses covering arguments in order to guarantee achievability. The result
is of the following flavor: there exists at least one $\Phi$ in ${\mathbb R}^{m \times n}$ that has RIP ratio equal to or better (smaller) than the one given by the
covering bound. The main result of this bound is Theorem \ref{theo:cswachievable} in Section \ref{sec:mainresults}. Again, our derivation considers only the case $k=2$;
we plan to extend the results for $k>2$ in a future publication.
After deriving the bounds, we present extensive numerical results from which we draw several important observations and conclusions.
A summary of our key results made visual in the Figures is given below:
\begin{enumerate}
\item We see from Figure \ref{fig-bounds-results} that there is a large gap between the RIP ratio of Gaussian matrices and the achievable
bounds. This observation points to the existence of CS matrices that are far superior than Gaussian
matrices in terms of the RIP ratio.
\item We see from Figure \ref{fig-bounds-results} that, for small values of $n$, the structural bound is the tighter of the two converse
bounds, whereas for large values of $n$, the packing bound is tighter.
\item Figure \ref{fig-bounds-results} also presents a graphical illustration of all three bounds and compares the bounds to the
RIP ratio of Gaussian matrices, as well as the best known matrix for the given problem size. We use
the results of Conway, Hardin, and Sloane \cite{SloaneGrassmannian,SloaneGrassmannianWeb,SloaneGrassmannianWebTable}, who have run extensive computer simulations
to extract the best known packings in Grassmannian spaces.
\item While the structural bound for the RIP ratio has been derived for any $n$, $m$, and $k$, we presently
have the packing and covering bounds only for $k=2$.
Figure~\ref{fig-strucbound} depicts the structural bound for $k=4$ and $k=10$.
\item Figure \ref{fig-histo} shows how the parameters of the structural bound can shed light on the statistical
RIP that was proposed recently~\cite{Tropp2008_1,Tropp2008_2,CalderbankDetRIP,gurevich}. In particular, we demonstrate a way to estimate the singular
values of a randomly chosen submatrix of $\Phi$ in Section~\ref{sec:stochrip}.
\end{enumerate}
Furthermore, we show that the structural bound for $k=2$ is equivalent to the Welch bound~\cite{WelchEquiangular}.
The structural bound for $k>3$ can be thought of as extensions to the Welch bound to higher orders $k>2$.
We state our extension to the Welch bound explicitly in Theorem~\ref{th:welchextension}.
The results we present offer lower bounds for the RIP ratio $R(\Phi, k)$ defined in (\ref{eq:ripratio}). The reason we do not derive the bounds directly for the Restricted Isometry constant $\delta_k$ defined in (\ref{eq:riconstant})
is because $\delta_k$ changes as we multiply $\Phi$ by a scalar. On the other hand, the RIP ratio
$R(\Phi,k)$ is invariant to the scaling of $\Phi$. We relate the lower bound on $R(\Phi, k)$ to lower bounds on $\delta_k$ using Theorem~\ref{th:bestripconstant} below.
First, fix $\Phi \in {\mathbb R}^{m \times n}$, $\Phi \ne 0$, and select the smallest $\epsilon_1 \in {\mathbb R}$ and $\epsilon_2 \in {\mathbb R}$ such that
\begin{equation}
(1-\epsilon_1) \|x\|_{\ell_2}^{2} \le \| \Phi x \|_{\ell_2}^{2} \le (1+\epsilon_2) \|x\|_{\ell_2}^{2}
\label{eq:eps12}
\end{equation}
holds for all $k$-sparse $x$.
From the definition of the Restricted Isometry constant $\delta_k$ in (\ref{eq:riconstant}),
we have $\delta_k = \max \{\epsilon_1, \epsilon_2\}$.
Now consider a scalar $a \in {\mathbb R}$, $a \ne 0$, and let $\Phi' = a\Phi$. select the smallest $\epsilon'_1 \in {\mathbb R}$ and $\epsilon'_2 \in {\mathbb R}$ such that
\begin{equation}
(1-\epsilon'_1) \|x\|_{\ell_2}^{2} \le \| \Phi' x \|_{\ell_2}^{2} \le (1+\epsilon'_2) \|x\|_{\ell_2}^{2}
\label{eq:eps212}
\end{equation}
holds for all $k$-sparse $x$. The Restricted Isometry constant for $\Phi'$, which we denote
by $\delta'_k$, is given by $\delta'_k = \max \{\epsilon'_1, \epsilon'_2\}$.
In this setting, we present the choice of the scalar $a$ that minimizes $\delta'_k$, i.e., we find the CS matrix that has the best Restricted Isometry constant
in the family of matrices $\Phi' = a \Phi$.
\begin{THEO}
Let $\Phi' = a \Phi$, where $a \in {\mathbb R}$, $a \ne 0$ is a scalar, and let $\epsilon_1$, $\epsilon_2$, $\epsilon'_1$, $\epsilon'_2$ be the smallest real numbers such that the
conditions of (\ref{eq:eps12}) and (\ref{eq:eps212}) are satisfied for $k$-sparse signals $x$, where $x \ne 0$. Let $\delta_k = \max \{\epsilon_1, \epsilon_2\}$ and
$\delta'_k = \max \{\epsilon'_1, \epsilon'_2\}$.
If
\begin{equation}
a = \frac{2}{2+\epsilon_2 - \epsilon_1},
\label{eq:aaa}
\end{equation}
then the following results hold:
\begin{enumerate}
\item $\delta'_k = \epsilon'_1 = \epsilon'_2$,
\item $\delta'_k \le \delta_k$.
\end{enumerate}
\label{th:bestripconstant}
\end{THEO}
{\bf Proof:}
We first note that
$\frac{1 - \epsilon'_1}{1 - \epsilon_1} = a$ and $\frac{1 + \epsilon'_2}{1 + \epsilon_2} = a$.
Using the choice of $a$ given in (\ref{eq:aaa}),
we express $\epsilon'_1$ and $\epsilon'_2$ in terms of $\epsilon_1$ and $\epsilon_2$ to obtain
\begin{equation*}
\epsilon'_1 = \frac{\epsilon_1 + \epsilon_2}{2 + \epsilon_2 - \epsilon_1} \mbox{ ~~~ and ~~~ } \epsilon'_2 = \frac{\epsilon_1 + \epsilon_2}{2 + \epsilon_2 - \epsilon_1}.
\end{equation*}
Therefore, $\epsilon'_1 = \epsilon'_2$. Since $\delta'_k = \max\{\epsilon'_1, \epsilon'_2\}$, we have $\delta'_k = \epsilon'_1 = \epsilon'_2$, proving
the first statement of the Theorem. \\
In order to prove the second statement of the Theorem, we consider three cases. \\
{\bf Case 1:} $\epsilon_1 = \epsilon_2$. \\
In this case, we have $a=1$, and so
$\epsilon'_1 = \epsilon_1$ and $\epsilon'_2 = \epsilon_2$. Therefore $\delta'_k = \delta_k$, which satisfies the
second statement of the Theorem. \\
{\bf Case 2:} $\epsilon_1 < \epsilon_2$. Under this assumption, $\delta_k = \epsilon_2$.\\
Because $\epsilon_1 < \epsilon_2$, we have $(2 - \epsilon_1 + \epsilon_2) > 2$, and so for the choice of $a$ in (\ref{eq:aaa}), we have
\begin{eqnarray}
1 + \epsilon'_2 &=& \frac{2}{2 + \epsilon_2 - \epsilon_1} (1 + \epsilon_2) \nonumber \\
&<& 1 + \epsilon_2. \nonumber
\end{eqnarray}
Therefore, we have $\delta'_k = \epsilon'_2 < \epsilon_2 = \delta_k$, which verifies the second statement of the Theorem. \\
{\bf Case 3:} $\epsilon_1 > \epsilon_2$. Under this assumption, $\delta_k = \epsilon_1$.\\
First we recognize that $(1 - \epsilon_1) \ge 0$ and $(1+\epsilon_2) > 0$, and therefore, the sum
$(1 - \epsilon_1) + (1+\epsilon_2) = (2 - \epsilon_1 + \epsilon_2) > 0$.
Since $\epsilon_1 > \epsilon_2$, we have $(2 - \epsilon_1 + \epsilon_2) < 2$. Hence, $0 < (2 - \epsilon_1 + \epsilon_2) < 2$.
For the choice of $a$ in (\ref{eq:aaa}), we have
\begin{eqnarray}
1 - \epsilon'_1 &=& \frac{2}{2 + \epsilon_2 - \epsilon_1} (1 - \epsilon_1) \nonumber \\
&>& 1 - \epsilon_1. \nonumber
\end{eqnarray}
Therefore, we have $\delta'_k = \epsilon'_1 < \epsilon_1 = \delta_k$. \\
Hence, we have $\delta'_k < \delta_k$ for all the 3 cases, proving the second statement of the Theorem. \qed
Theorem~\ref{th:bestripconstant} shows how we can pick $a$ in order to obtain the matrix with the best Restricted Isometry constant
(namely $\delta'_k$)
from the family of matrices $a\Phi$. Below, we express $\delta_k'$ in terms of the RIP ratio, which is scale invariant.
Recognizing that $1-\epsilon_1 = \rho_{\min}(\Phi,k)$ and $1+\epsilon_2 = \rho_{\max}(\Phi,k)$, we can write $\delta'_k$ as
\begin{eqnarray}
\delta'_k &=& \frac{\epsilon_2+\epsilon_1}{2+\epsilon_2-\epsilon_1}
\mbox{~~~~~~~~~~~~~~~~~~(from Theorem~\ref{th:bestripconstant})} \nonumber \\
&=& \frac{(1+\epsilon_2)-(1-\epsilon_1)}{(1+\epsilon_2)+(1-\epsilon_1)} \nonumber \\
&=& \frac{\rho_{\max}(\Phi , k) - \rho_{\min}(\Phi , k)}{\rho_{\max}(\Phi , k) + \rho_{\min}(\Phi , k)} \nonumber \\
&=& \frac{R(\Phi,k) - 1}{R(\Phi,k) + 1}. \label{eq:deltakripratioequiv}
\end{eqnarray}
Using (\ref{eq:deltakripratioequiv}), we infer that the condition $\delta_{2k} < \sqrt{2}-1$ in Theorems~\ref{th:l0l1equal} and \ref{th:noisycs}
is equivalent to the condition $R(\Phi , 2k) < \sqrt{2}+1$. More precisely, if $\Phi$ satisfies
$R(\Phi , 2k) < \sqrt{2}+1$, then there exists a scalar $\alpha \ne 0$ such that the scaled matrix $\alpha \Phi$ satisfies $\delta_{2k} < \sqrt{2}-1$.
\subsection{Organization of the Paper}
In
Section~\ref{sec:mainresults},
we present the three main Theorems (Theorems~\ref{theo:main1}, \ref{theo:packing} and \ref{theo:cswachievable})
with the three deterministic bounds on the RIP ratio. We also present a graphical illustration of the
three bounds in comparison with RIP ratios of real matrices (including Gaussian matrices) taken from ${\mathbb R}^{m \times n}$.
In Section~\ref{sec:structuralbound} we prove the structural bound for the RIP ratio and present
numerous results on the properties of this bound.
We also offer a tighter structural bound (Theorem \ref{theo:thompsonrip}) when the singular values of $\Phi$ are known. We end the Section by offering
a geometric interpretation of the bounds and their relationship to the Generalized Pythagorean Theorem and to equi-angular tight frames (ETF).
In Section~\ref{sec:packing}, we show the equivalence of optimizing the RIP ratio of CS matrices and optimal packing in Grassmannian spaces. We use packing and covering arguments to
prove the deterministic packing bound and the achievable bound.
In Section~\ref{sec:stochrip},
we demonstrate how one can extend the results on
deterministic bounds to statistical-RIP. We conclude in Section~\ref{sec:conclusions}.
\section{Main Results}
\label{sec:mainresults}
\subsection{Structural Bound: Converse Deterministic Bound for RIP Ratio}
\begin{THEO} {\em (Structural Bound)}
Let $\Phi$ be an ${m \times n}$ matrix over ${\mathbb R}$ or ${\mathbb C}$ and $0 < m <n$.
Let $k$ be an integer such that $0 < k <m$ and define the $k$'th degree polynomial
\begin{equation}
f_k(x) \triangleq D^{n-k} \left[ x^{n-m} (x-1)^m \right].
\label{eq:fkx}
\end{equation}
The following results are true:
\begin{enumerate}
\item The $k$ zeros of $f_k(x)$ are real and lie in the interval $(0,1]$.
\item Let $r_1^2 \ge r_2^2 \ge ... r_k^2$ be the zeros of $f_k(x)$.
Then, the following lower bound on the RIP ratio holds
\begin{equation}
R(\Phi, k) \ge \frac{r_1^2}{r_k^2}.
\label{eq:RIPbound1}
\end{equation}
\item Let $S_1 \ge S_2 \ge ... \ge S_m > 0$ be the $m$ singular values of $\Phi$.
Equality in (\ref{eq:RIPbound1})
is achieved if and only if all of the following three conditions are satisfied:
\begin{enumerate}
\item $S_1 = S_2 = ... = S_m$.
\item The largest singular values of every $m \times k$ submatrix of $\Phi$
are all equal.
That is, $s_{p,1}=s_{q,1}$ for all $p, q \in \{1,2,...,{n \choose k} \}$.
\item The smallest singular values of every $m \times k$ submatrix of $\Phi$
are all equal.
That is, $s_{p,k}=s_{q,k}$ for all $p, q \in \{1,2,...,{n \choose k} \}$.
\end{enumerate}
\end{enumerate}
\label{theo:main1}
\end{THEO}
The above Theorem asserts that we cannot find a $\Phi$ in ${\mathbb R}^{m \times n}$ that has a
better (smaller) RIP ratio $R(\Phi , k)$ than $r_1^2/r_k^2$.
\subsection{Packing Bound: Converse Deterministic Bound for RIP Ratio for $k=2$}
\begin{DEFI}
Define the function $c_m(\beta)$, $\beta \in [0, \pi]$ by
\begin{equation*}
c_m(\beta) = \int_0^\beta \sin^{m-2} \alpha d\alpha.
\end{equation*}
\end{DEFI}
\begin{THEO} {\em (Packing Bound)}
Let $\theta \in (0,\pi)$ be the solution to the equation
\begin{equation}
c_m\left( \theta \right) = \frac{c_m(\pi)}{n},
\label{eq:theta}
\end{equation}
and let
$q_1 = \cot^2 \left( \theta \right)$.
Then there exists no $\Phi \in {\mathbb R}^{m \times n}$ that satisfies
$R(\Phi,2) \le q_1$.
\label{theo:packing}
\end{THEO}
The above Theorem asserts that we cannot find a $\Phi$ in ${\mathbb R}^{m \times n}$ that has a
better (smaller) RIP ratio $R(\Phi , 2)$ than $q_1$.
Note that because $c_m(\beta)$ increases monotonically with $\beta \in (0, \pi)$, the solution to (\ref{eq:theta}) can easily be solved numerically.
\subsection{Covering Bound: Achievable Deterministic Bound for RIP Ratio for $k=2$}
\begin{THEO} {\em (Covering Bound)}
Let $\theta \in (0,\pi)$ be the solution to the equation
\begin{equation*}
c_m(\theta) = \frac{c_m(\pi)}{n},
\end{equation*}
and let
$q_2 = \cot^2 \left( \frac{\theta}{2} \right)$.
Then, there exists a $\Phi \in {\mathbb R}^{m \times n}$ such that
$R(\Phi,2) \le q_2$.
\label{theo:cswachievable}
\end{THEO}
The above Theorem
guarantees the existence of a
$\Phi$ in ${\mathbb R}^{m \times n}$ that has a
better (smaller) RIP ratio $R(\Phi , 2)$ than $q_2$.
\subsection{Graphical Illustration of the Bounds for $k=2$}
Figure \ref{fig-bounds-results} illustrates the three bounds presented above, along with
the performance of a Gaussian CS matrix and the ``best'' known CS matrix. The results are given for $k=2$, for which we
can compute all the three bounds presented above. Furthermore,
for $k=2$, we show in Section~\ref{sec:packing} that finding the best CS matrix (in terms of RIP ratio) is equivalent to a coding problem in Grassmannian spaces. Fortunately,
we have the best known codes available from the work of Conway, Hardin, and Sloane \cite{SloaneGrassmannian,SloaneGrassmannianWeb,SloaneGrassmannianWebTable}; therefore, we can compute the best CS matrices for a given choice of $n$ and $m$.
Hence we can compare the three bounds with the best known CS matrix as well as provide comparisons to a randomly generated CS matrix using i.i.d. Gaussian distribution.
In Figure \ref{fig-bounds-results}, we consider respectively $m=3$, $m=6$, $m=8$, $m=10$, $m=12$, and $m=16$.
We vary $n$ along the horizontal axis and plot the RIP ratio and the bounds on the vertical axis.
\begin{figure*}
\centering
\subfigure[$m= 3$] {\epsfysize = 58mm \epsffile{thompplots_m3.eps} \label{fig-subfig_m3}}
\subfigure[$m= 6$] {\epsfysize = 58mm \epsffile{thompplots_m6.eps} \label{fig-subfig_m6}} \\
\subfigure[$m= 8$] {\epsfysize = 58mm \epsffile{thompplots_m8.eps} \label{fig-subfig_m8}}
\subfigure[$m=10$] {\epsfysize = 58mm \epsffile{thompplots_m10.eps} \label{fig-subfig_m10}} \\
\subfigure[$m=12$] {\epsfysize = 58mm \epsffile{thompplots_m12.eps} \label{fig-subfig_m12}}
\subfigure[$m=16$] {\epsfysize = 58mm \epsffile{thompplots_m16.eps} \label{fig-subfig_m16}}
\caption{\sl RIP ratio bounds for $k=2$ and different values of $m$ and $n$. We compare the three bounds to the
RIP ratio of Gaussian matrices as well as the best known matrix for the given problem size. The RIP ratio of
Gaussian $\Phi$ shown is the geometric mean of the RIP ratio of 10,000 instantiations of randomly generated $m \times n$
Gaussian matrices.
The RIP ratio of the best known $\Phi$ is based on
the results of Conway, Hardin, and Sloane~\cite{SloaneGrassmannian,SloaneGrassmannianWeb,SloaneGrassmannianWebTable} on the best known packings in Grassmannian spaces.}
\label{fig-bounds-results}
\end{figure*}
The following observations can be made from Figure~\ref{fig-bounds-results}.
\begin{enumerate}
\item The Gaussian CS matrix construction has a much higher RIP ratio in comparison to the three bounds, revealing the existence of CS matrices in
${\mathbb R}^{m \times n}$ that will offer much better performance. Specifically, there is a substantial
gap between the Gaussian RIP ratio and the
achievable bound.
\item The best known CS matrix (obtained from the work of Conway, Hardin and Sloane~\cite{SloaneGrassmannian,SloaneGrassmannianWeb,SloaneGrassmannianWebTable})
in ${\mathbb R}^{m \times n}$ has an RIP ratio that is lower than the covering bound and greater than the structural and packing bounds. Therefore, we can think of the
covering bound as an upper bound and the structural and packing bounds as lower bounds; together, they define a band within
which the RIP ratio of the optimal CS matrix resides. Note that the best {\em known} $\Phi$ lies within this band.
In fact, the RIP ratio curve for the best $\Phi$ performs significantly better than the achievable bound.
\item Regarding the converse bound, the structural bound is stronger for smaller values of $n$, whereas the packing bound is stronger for larger values of $n$.
In fact, for $n < m(m+1) /2$, the best CS matrices are governed by the structural bound, whereas for $n > m(m+1) / 2$, the packing bound governs the behavior.\footnote{For $m=16$,
the structural bound seems to govern for higher $n$ as well.}
In Section~\ref{sec:structuralbound}, we see the significance of this transition point in $n$ and relate it to equi-angular tight frames.
\end{enumerate}
\section{Structural Bound for the RIP Ratio}
\label{sec:structuralbound}
In this Section, we prove Theorem~\ref{theo:main1} by first presenting a series of
results. We also uncover the properties of the proposed lower bound on the RIP ratio.
\subsection{Bound for RIP Ratio when Singular Values of $\Phi$ are Known}
\label{sec:substructuralbound}
First, we
present a lower bound on the RIP ratio when the singular values of $\Phi$ are {\em known}.
\begin{THEO}
Let $\Phi$ be an ${m \times n}$ matrix over ${\mathbb R}$ or ${\mathbb C}$ and $0 < m <n$.
Let $S_1 \ge S_2 \ge ... \ge S_m $ be the $m$ singular values of $\Phi$.
Let the singular values of the submatrix $\Phi_p$ be
$s_{p,1} \ge s_{p,2} \ge...\ge s_{p,k}$, for all $p \in \{ 1,2,...,{n \choose k}\}$.
Let
\begin{equation}
g_k(x) = D^{n-k} \left[ x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2) \right].
\label{eq:thompson_result}
\end{equation}
Let $r_1^2 \ge r_2^2 \ge ... r_k^2$ be the zeros of $g_k(x)$.
Then, the following results are true:
\begin{enumerate}
\item The $k$ zeros of $g_k(x)$ are real and lie in the interval $(0,S_1^2]$.
\item $\rho_{\max}(\Phi,k) \ge r_1^2, ~~ \rho_{\min}(\Phi,k) \le r_k^2, ~~\mbox{ and }~~R(\Phi,k) \ge \frac{r_1^2}{r_k^2}.$
\item Equality in all three inequalities above is attained if and only if
the following two conditions are satisfied:
\begin{enumerate}
\item The largest singular values of every $m \times k$ submatrix of $\Phi$
are all equal.
That is, $s_{p,1}=s_{q,1}$ for all $p, q \in \{1,2,...,{n \choose k} \}$.
\item The smallest singular values of every $m \times k$ submatrix of $\Phi$ are all equal.
That is, $s_{p,k}=s_{q,k}$ for all $p, q \in \{1,2,...,{n \choose k} \}$.
\end{enumerate}
\end{enumerate}
\label{theo:thompsonrip}
\end{THEO}
The proof of this result hinges on a
Theorem of Robert C. Thompson~\cite{RCThompson9}
that relates the singular values of all submatrices of a given size to the
singular values of the complete matrix.
We present the statement of Thompson's Theorem here for completeness; its proof
and other results on singular~values and eigen~values of submatrices are dealt with in
~\cite{RCThompson1,RCThompson2,RCThompson3,RCThompson4,RCThompson5,RCThompson6,RCThompson7,RCThompson8,RCThompson9}.
\begin{THEO} {\em \cite[Theorem 4]{RCThompson9}}
Let $A$ be an $m \times n$ matrix with singular values $\alpha_1 \ge \alpha_2 \ge ... \ge \alpha_{\min\{m,n\}}$.
Let $B_p$ be an $l \times k$ submatrix of $A$ for some fixed $l \in \{1,2,...,m \}$ and $k \in \{1,2,...,n \}$ with
singular values $\beta_{p,1} \ge \beta_{p,2} \ge ... \ge \beta_{p,{\min\{l,k\}}}$, and index $p$
that identifies the submatrix amongst all $l \times k$ submatrices of $A$.
Set
\begin{eqnarray}
f_p(x) &=& (x-\beta_{p,1}^2)(x-\beta_{p,2}^2)...(x-\beta_{p,{\min\{l,k\}}}^2), \nonumber \\
f(x) &=& (x-\alpha_{1}^2)(x-\alpha_{2}^2)...(x-\alpha_{{\min\{m,n\}}}^2). \nonumber
\end{eqnarray}
Then, the following result is true:
\begin{equation*}
\sum_p x^{l-\min\{l,k\}} f_p(x) =
\frac{1}{(m-l)!}
\frac{1}{(n-k)!}
D^{m-l}\left[
x^{m-k}
D^{n-k}\left[
x^{n-\min\{ m,m\}} f(x)
\right]
\right],
\end{equation*}
where the summation is taken over all the submatrices of $A$ of size $l \times k$.
\label{theo:thompson_original}
\end{THEO}
\noindent {\bf Proof of Theorem~\ref{theo:thompsonrip}: }
The zeros of $g_k(x)$ are real and lie in the closed interval $[0, S_1^2]$
as a consequence of the Gauss-Lucas Theorem~\cite{Rahman2002}. Recall that the Gauss-Lucas
Theorem asserts that every convex set in the complex plane containing all the zeros of a polynomial also contains all its
critical points (the zeros of the derivative of the chosen polynomial). In our setting, we consider the polynomial
$G(x)=x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2)$ that has all its zeros in the closed interval $[0,S_1^2]$ on the real number line.
Differentiating the polynomial $G(x)$ $(n-k)$ times and applying the Gauss-Lucas Theorem at each step, we infer that
the $k$ zeros of $g_k(x)$ are real and lie in the interval $[0,S_1^2]$. Finally, we observe that $x=0$ cannot be a zero of the polynomial
$g_k(x)$, because $x=0$ is a zero of $G(x)$ of order $(n-m)$, whereas we differentiate $(n-k) > (n-m)$ times. Thus we have
established the first statement of Theorem~\ref{theo:thompsonrip}.\footnote{Additionally, note that when $n-k > m$,
$x=S_1^2$ cannot be a zero of $g_k(x)$.}
To prove the second statement of Theorem~\ref{theo:thompsonrip}, we apply Theorem~\ref{theo:thompson_original} (Thompson's Theorem) to our setting.
Consider the complete matrix $\Phi$ of size ${m \times n}$ and the collection of ${m \times k}$ sized
submatrices $\Phi_p$ for
$p=1,2,...,{n \choose k}$. We obtain
\begin{equation}
\frac{1}{(n-k)!}D^{n-k } \left[ x^{n-m}
(x-S_1^2)(x-S_2^2)...(x-S_m^2) \right]
=
\sum_{p} (x-s_{p,1}^2)(x-s_{p,2}^2)...(x-s_{p,k}^2).
\label{eq:thompson_basic}
\end{equation}
The above equation relates the singular value polynomial of $\Phi$ to the singular value
polynomials of the submatrices $\Phi_p$. Recall that the singular value polynomial of a matrix $A$
is the polynomial whose roots are the squares of the singular values of the matrix.\footnote{Equivalently,
the singular value polynomial of $A$ is the characteristic polynomial of the Grammian $A^HA$ or $AA^H$, whichever has
the higher degree.} Equation (\ref{eq:thompson_basic}) asserts that the sum of the singular value
polynomials of the submatrices $\Phi_p$ is a constant multiple of the $(n-k)$'th derivative of the
singular value polynomial of $\Phi$.
Since $r_1^2, r_2^2,...,r_k^2$ are the roots of $g_k(x)$, we have
\begin{equation}
\frac{n!}{k!} (x-r_1^2)(x-r_2^2)...(x-r_k^2) = D^{n-k} \left[ x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2) \right],
\label{eq:thompson_roots}
\end{equation}
where the constant $n!/k!$ is needed to equalize the coefficients of $x^k$ in the LHS and RHS.
From (\ref{eq:thompson_basic}) and (\ref{eq:thompson_roots}), we have
\begin{equation}
{n \choose k} (x-r_1^2)(x-r_2^2)...(x-r_k^2)
=
\sum_{p} (x-s_{p,1}^2)(x-s_{p,2}^2)...(x-s_{p,k}^2).
\label{eq:thompson_rs}
\end{equation}
We are now in a position to prove the second statement of Theorem~\ref{theo:thompsonrip}.
First, we note that $\rho_{\max}(\Phi,k) = \max_p \{ s_{p,1}^2 \}$ and $\rho_{\min}(\Phi,k) = \min_p \{ s_{p,1}^2 \}$.
The first inequality in the second statement implies that $\max_p \{s_{p,1}^2\} \ge r_1^2$.
For the sake of a contradiction, assume that $s_{p,1}^2 < r_1^2$ for all $p \in \left\{1,2,... {n \choose k} \right\}$.
Under this assumption, $(r_1^2-s_{p,i}^2) > 0 $ for all $p$ and $i$ and hence each of the polynomial $(x-s_{p,1}^2)(x-s_{p,2}^2)...(x-s_{p,k}^2)$
is strictly positive when evaluated at $x=r_1^2$. Consequently, the sum in the RHS of (\ref{eq:thompson_rs})
evaluated at $x=r_1^2$ is strictly positive, which is a contradiction, because the LHS of (\ref{eq:thompson_rs})
evaluates to zero for $x=r_1^2$.
Therefore, we require $r_1^2 \le \max \{s_{p,1}^2\} = \rho_{\max}(\Phi,k)$.
We can show that $r_k^2 \ge \max \{s_{p,k}^2\} = \rho_{\min}(\Phi,k)$ using a similar argument as above. Note that
we need to consider the additional nuance of the sign of $(x-s_{p,1}^2)(x-s_{p,2}^2)...(x-s_{p,k}^2)$
when evaluating at $x=r_k^2$ with the assumption $s_{p,k}^2 > r_k^2$.
The sign of $(x-s_{p,1}^2)(x-s_{p,2}^2)...(x-s_{p,k}^2)$ is either positive or negative depending on whether
$k$ is even or odd, respectively.
Finally, since $\rho_{\max}(\Phi,k) \ge r_1^2$ and $\rho_{\min}(\Phi,k) \le r_k^2$, we have $R(\Phi,k) \ge \frac{r_1^2}{r_k^2}$, and hence
the second statement of Theorem~\ref{theo:thompsonrip} is established.
We now prove the third statement of Theorem~\ref{theo:thompsonrip}.
When $s_{p,1}^2$ are equal for all $p \in \{1,2,...,{n \choose k} \}$, we observe that with $x=s_{p,1}^2$, the RHS of
(\ref{eq:thompson_rs}) evaluates to zero. Hence $s_{p,1}^2$ must equal $r_i^2$ for
some $i \in \{1,2,...,k \}$. Since we have established in the second statement of Theorem~\ref{theo:thompsonrip} that we cannot have a zero
of $(x-r_1^2)(x-r_2^2)...(x-r_k^2)$ greater than $s_{p,1}^2$, it follows that $r_1^2=s_{p,1}^2$ for all $p$.
Similarly, when $s_{p,k}^2$ are equal for all possible $p$, we have $r_{k}^2 = s_{p,k}^2$.
Conversely, suppose $r_1^2 = \max\{ s_{p,1}^2\}$ is true. For the sake of a contradiction assume that the
values of $s_{p,1}^2$, $p \in \{1,2,...,{n \choose k} \}$ are not all equal. In particular, pick $p_1$ such that $s_{p_1,1}^2 < \max\{ s_{p,1}^2\}$.
Then the polynomial corresponding to $p=p_1$ within the summation in RHS in (\ref{eq:thompson_rs}) is strictly
positive when evaluated at $r_1^2$. Consequently, $r_1^2$ cannot be a zero of (\ref{eq:thompson_rs}),
yielding a contradiction.
A similar argument applies to the case when $r_k^2 = \min\{ s_{p,k}^2\}$.
Thus we have established the third statement of Theorem~\ref{theo:thompsonrip}, and therefore the proof of
Theorem~\ref{theo:thompsonrip} is complete.
\qed
While Theorem~\ref{theo:thompsonrip} applies to the case where the singular values of the
CS matrix $\Phi$ are known, Theorem~\ref{theo:main1} applies to {\em all} matrices $\Phi$ of
size $m \times n$. In our quest for deterministic bounds for the RIP ratio, Theorem~\ref{theo:main1}
therefore is of central importance.
To establish Theorem~\ref{theo:main1}, we explore the following question: what choice of
$S_1^2, S_2^2,...,S_m^2$ gives the most conservative bound for the RIP ratio
when we invoke Theorem~\ref{theo:thompsonrip}? We show below
(Theorem~\ref{theo:mincondition}) that the ratio $\frac{r_1^2}{r_k^2}$ is minimized when all the
singular values of $\Phi$ are equal. Therefore this minimum ratio serves as a universal bound for RIP
applicable to {\em all} CS matrices $\Phi$ of size $m \times n$ and leads to the proof
of Theorem~\ref{theo:main1}.\footnote{Of course, Theorem~\ref{theo:thompsonrip}
provides a tighter bound for the RIP ratio when the singular values of $\Phi$ are known.}
Toward this goal,
we investigate the nature of the dependence of $r_i^2$ on $S_1^2,S_2^2,...,S_m^2$ in Section~\ref{sec:partialdiff}.
\subsection{The Zeros of $g_k(x)$ and the Singular Values of $\Phi$}
\label{sec:partialdiff}
In this Section, we treat each $r_i^2$, $i \in \{1,2,...,k\}$ as a function of $m$ variables $S_1^2,S_2^2,...,S_m^2$ with the goal to determine the optimal choice of $S_j$'s that minimize the ratio $\frac{r_1^2}{r_k^2}$. The exact functional form is
given in (\ref{eq:thompson_roots}).
First, we present how an infinitesimal increment in one of the $S_j^2$'s affects the values of
$r_i^2$'s.
\begin{THEO}
Let $G(x)=x^{n-m}(x-S_1^2)(x-S_2^2)...(x-S_m^2)$ and let
$r_1^2 \ge r_2^2 \ge ... \ge r_k^2$ be the zeros of the polynomial
$g_k(x)=D^{n-k}[G(x)]$. Then, we have
\begin{eqnarray}
\frac{\partial (r_i^2)}{ \partial (S_j^2)} &=& \left[ \frac{D^{n-k} \left( \frac{G(x)}{x-S_j^2} \right)}{D^{n-k+1} G(x)} \right] ~~ \mbox{evaluated at} ~~ {x=r_i^2} \label{eq:partialdiff1} \nonumber \\
&=& \left[x \frac{ D^{n-k} \left( \frac{G(x)}{x-S_j^2} \right)}
{ \sum_{l=1}^{m} S_l^2 D^{n-k} \left( \frac{G(x)}{x-S_l^2}\right) } \right] ~~ \mbox{evaluated at} ~~ {x=r_i^2}. \label{eq:partialdiff2}
\end{eqnarray}
\label{theo:partialdiff}
\end{THEO}
\noindent {\bf Proof:} Taking the partial derivative of (\ref{eq:thompson_roots}) with respect to the
variable $S_1^2$ while keeping $S_2^2, S_3^2, ... , S_m^2$ fixed, we have
\begin{eqnarray}
\frac{n!}{k!} \frac{\partial}{\partial (S_1^2)}(x-r_1^2)...(x-r_k^2) = \frac{\partial}{\partial (S_1^2)} D^{n-k} \left[ x^{n-m} (x-S_1^2)...(x-S_m^2) \right].
\label{eq:partial1}
\end{eqnarray}
Treating the quantity inside the $D^{n-k}[\cdot]$ as a product of $(x-S_1^2)$ and $x^{n-m} (x-S_2^2)(x-S_3^2)...(x-S_m^2)$ and
using Leibnitz's Theorem~\cite{CalculusAlbert} for the $(n-k)$'th derivative of a product, we obtain
\begin{eqnarray}
D^{n-k} \left[ x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2) \right]
= (x-S_1^2) D^{n-k}\left[ x^{n-m} (x-S_2^2)(x-S_3^2)...(x-S_m^2) \right] \nonumber \\
+ (n-k)D^{n-k-1}\left[ x^{n-m} (x-S_2^2)(x-S_3^2)...(x-S_m^2) \right].
\label{eq:partial2}
\end{eqnarray}
Taking partial derivative with respect to $S_1^2$ (and noting that the second term of RHS in (\ref{eq:partial2})
is independent of $S_1^2$), we can rewrite the RHS of (\ref{eq:partial1}) as
\begin{eqnarray}
\frac{\partial}{\partial (S_1^2)} D^{n-k} \left[ x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2) \right] & & \nonumber
\end{eqnarray}
\begin{eqnarray}
&=& - D^{n-k}\left[ x^{n-m} (x-S_2^2)(x-S_3^2)...(x-S_m^2) \right] \nonumber \\
&=& -D^{n-k} \left[ \frac{G(x)}{(x-S_1^2)} \right].
\label{eq:partial3}
\end{eqnarray}
Expanding the LHS of (\ref{eq:partial1}),
\begin{eqnarray}
\frac{n!}{k!}\frac{\partial}{\partial (S_1^2)}(x-r_1^2)...(x-r_k^2)
&=& - \frac{n!}{k!}\sum_{i=1}^{k} \left\{ \left[ \prod_{1\le l \le k, l \ne i}(x-r_l^2) \right]\frac{\partial r_i^2}{\partial (S_1^2)} \right\}.
\label{eq:partial4}
\end{eqnarray}
Evaluating (\ref{eq:partial4}) at $x=r_1^2$ and noting that only one term in the summation in the RHS of
(\ref{eq:partial4}) is non-zero, we obtain
\begin{eqnarray}
\frac{n!}{k!} \left. \frac{\partial}{\partial (S_1^2)}(x-r_1^2)...(x-r_k^2) \right|_{x=r_i^2}
&=& - \left. \frac{n!}{k!} \left[ \prod_{1\le l \le k, l \ne i}(x-r_l^2) \right] \frac{\partial r_i^2}{\partial (S_1^2)} \right|_{x=r_i^2} \nonumber \\
&=& - \left. \frac{n!}{k!} \frac{\left[ \prod_{1\le l \le k}(x-r_l^2) \right]}{x-r_1^2}\frac{\partial r_i^2}{\partial (S_1^2)} \right|_{x=r_i^2}\nonumber \\
&=& - \left. \frac{D^{n-k}\left[ G(x) \right]}{x-r_1^2} \frac{\partial (r_1^2)}{\partial (S_1^2)} \right|_{x=r_i^2}.
\label{eq:partial5}
\end{eqnarray}
From (\ref{eq:partial1}), (\ref{eq:partial3}), and (\ref{eq:partial5}), we infer that
\begin{eqnarray}
\frac{\partial (r_i^2)}{ \partial (S_j^2)} &=&
\left[ \frac{D^{n-k} \left( \frac{G(x)}{x-S_j^2} \right)}{\frac{D^{n-k} G(x)}{x-r_1^2} } \right] ~~ \mbox{evaluated at} ~~ {x=r_i^2}.
\label{eq:partial7}
\end{eqnarray}
Finally, we use the well known result that if a function $f(x)$ has a zero at $x=a$ of multiplicity 1 (i.e., a simple zero),
then $f(x)/(x-a)$ evaluated at $x=a$ is equal to $D\left[ f(x) \right]$ evaluated at $x=a$. Applying this result, we have that
$D^{n-k} G(x)/(x-r_1^2)$ evaluated at $x=r_1^2$ is equal to $D^{n-k+1} G(x)$ evaluated at $x=r_1^2$, since $r_1^2$ is a
zero of $D^{n-k} G(x)$. Using this in (\ref{eq:partial7}), we deduce that
\begin{eqnarray}
\frac{\partial (r_i^2)}{ \partial (S_j^2)} &=&
\left[ \frac{D^{n-k} \left( \frac{G(x)}{x-S_j^2} \right)}{D^{n-k+1} G(x)} \right] ~~ \mbox{evaluated at} ~~ {x=r_i^2},
\label{eq:partial8}
\end{eqnarray}
which proves the first part of Theorem~\ref{theo:partialdiff}.
To prove the second part, denote $S(x)=(x-S_1^2)(x-S_2^2)...(x-S_m^2)$, and consider the quantity $D^{n-k+1}[xG(x)]$. We have
\begin{eqnarray}
D^{n-k+1}[xG(x)] &=& D^{n-k} \left[ D\left\{ x^{n-m+1} S(x) \right\} \right] \nonumber \\
&=& D^{n-k}\left[ (n-m+1)x^{n-m} S(x) + x^{n-m+1} D[S(x)] \right] \nonumber \\
&=& (n-m+1) D^{n-k} \left[x^{n-m}S(x)\right] + D^{n-k}\left[ x D[S(x)] \right].
\label{eq:partial11}
\end{eqnarray}
Alternately, we can use Leibnitz Theorem to expand $D^{n-k+1}[xG(x)]$, yielding
\begin{eqnarray}
D^{n-k+1}[xG(x)] &=& (n-k+1)D^{n-k} \left[ G(x) \right] + xD^{n-k+1} \left[ x^{n-m} S(x) \right].
\label{eq:partial12}
\end{eqnarray}
From (\ref{eq:partial11}) and (\ref{eq:partial12}), we have
\begin{eqnarray}
(n-m+1) D^{n-k} \left[x^{n-m}S(x)\right] &+& D^{n-k}\left[ x D[S(x)] \right] \nonumber \\
=~~~ (n-k+1)D^{n-k} \left[ G(x) \right] &+& xD^{n-k+1} \left[ x^{n-m} S(x) \right]
\label{eq:partial13}
\end{eqnarray}
and therefore
\begin{eqnarray}
D^{n-k}\left[ x^{n-m+1} D[S(x)] \right] &=& (m-k)D^{n-k} \left[ G(x) \right] + xD^{n-k+1} \left[ G(x) \right].
\label{eq:partial14}
\end{eqnarray}
Therefore,
\begin{eqnarray}
D^{n-k+1} \left[ x^{n-m} S(x) \right] &=& \frac{1}{x}\left\{ D^{n-k}\left[ x^{n-m+1} D[S(x)] \right] - (m-k)D^{n-k} \left[ G(x) \right] \right\} \nonumber \\
&=& \frac{1}{x}\left\{ D^{n-k}\left[ x^{n-m+1} D[S(x)] -m G(x) \right] + k D^{n-k} \left[ G(x) \right] \right\}. \label{eq:partial15}
\end{eqnarray}
Since $D[S(x)]=S(x) \left\{ \frac{1}{x-S_1^2} + ... + \frac{1}{x-S_m^2} \right\}$, we have
\begin{eqnarray}
x^{n-m+1}D[S(x)] &=& x^{n-m} S(x) \left\{ \frac{x}{x-S_1^2} + ... + \frac{x}{x-S_m^2} \right\}.
\label{eq:partial16}
\end{eqnarray}
Rewriting $m D^{n-k} G(x)$ as
\begin{eqnarray}
m D^{n-k} G(x)
&=& D^{n-k}\left[ x^{n-m}S(x) \left\{ \frac{x-S_1^2}{x-S_1^2} + ... + \frac{x-S_m^2}{x-S_m^2} \right\} \right]
\label{eq:partial17}
\end{eqnarray}
and, subtracting (\ref{eq:partial17}) from (\ref{eq:partial16}), we obtain
\begin{eqnarray}
D^{n-k}\left[ x^{n-m+1} D[S(x)] \right] &-& m D^{n-k} G(x) \nonumber \\
&=& ~~ D^{n-k}\left[ x^{n-m}S(x) \left\{ \frac{S_1^2}{x-S_1^2} + ... + \frac{S_m^2}{x-S_m^2} \right\} \right].
\label{eq:partial18}
\end{eqnarray}
Plugging the above result in (\ref{eq:partial15}) yields
\begin{eqnarray}
D^{n-k+1} \left[ x^{n-m} S(x) \right] &=& \frac{1}{x} k D^{n-k} \left[ G(x) \right] \nonumber \\
&+& \frac{1}{x} \left\{ \sum_{l=1}^{m} S_l^2 D^{n-k} \left( \frac{G(x)}{x-S_l^2}\right) \right\} .
\label{eq:partial21}
\end{eqnarray}
Substituting (\ref{eq:partial21}) in (\ref{eq:partial8}) and evaluating at $x=r_1^2$ (noting that $D^{n-k} \left[ G(x) \right]$ is zero at $x=r_1^2$), we obtain
the result in (\ref{eq:partialdiff2}) which completes the proof of Theorem~\ref{theo:partialdiff}.
\qed
\begin{REMA}
From (\ref{eq:partialdiff2}), we have
\begin{equation}
S_1 \frac{\partial (r_i^2)}{ \partial (S_1^2)} + S_2 \frac{\partial (r_i^2)}{ \partial (S_2^2)} + ... + S_m \frac{\partial (r_i^2)}{ \partial (S_m^2)} = r_i^2,
\label{eq:eulerhomogeneous}
\end{equation}
which we recognize as Euler's condition for homogeneity of $r_i^2$ on $\{S_1^2, S_2^2, ... , S_m^2 \}$ of degree $1$~\cite{CalculusAlbert}.
\label{rema:euler}
\end{REMA}
Remark~\ref{rema:euler} implies that
when we write $r_i^2$ as a function of $\{S_1^2, S_2^2, ... , S_m^2 \}$, we have
\begin{equation}
r_i^2(aS_1^2, aS_2^2,...,aS_m^2) = a r_i^2(S_1^2, S_2^2,...,S_m^2),
\label{eq:homogenouscondition}
\end{equation}
for all $a \in {\mathbb R}^{+}$. This result is in agreement with (\ref{eq:thompson_roots}), because the homogeneity result can be
derived from (\ref{eq:thompson_roots}) by making a change of variable from $x$ to $ax$.
\begin{THEO}
Under the assumptions of Theorem~\ref{theo:partialdiff},
\begin{equation*}
\frac{\partial (r_i^2)}{ \partial (S_j^2)} \ge 0.
\end{equation*}
\label{theo:rige0}
\end{THEO}
Theorem~\ref{theo:rige0} is significant, because it indicates that increasing $S_j^2$ for any $j$ can only increase
the corresponding $r_i^2$'s. This fact can be exploited if our objective is to maximize the
singular values of the submatrices of $\Phi$. In Donoho's paper ~\cite{DonohoCS},
the condition CS1 requires the smallest singular values of the submatrices of $\Phi$ to exceed
a positive constant. The above Theorem indicates the relationship of CS1 condition to the
singular values of the complete matrix $\Phi$.
Also, in any practical system, we have a bound on the maximum $S_j^2$'s that can be used, reminiscent of
coding with power constraints~\cite{DigiComm}. In such a scenario, Theorem~\ref{theo:rige0} indicates that
the best choice for the $S_j^2$'s is when they are all equal to the maximum allowable bound.
In order to prove Theorem~\ref{theo:rige0}, we require a result on interlacing polynomials.
\begin{DEFI}
Two non-constant polynomials $p(x)$ and $q(x)$ with real coefficients have {\em weakly interlacing zeros} if:
\begin{itemize}
\item their degrees are equal or differ by one,
\item their zeros are all real, and
\item there exists an ordering such that
\begin{equation}
\alpha_1 \le \beta_1 \le \alpha_2 \le \beta_2 \le ... \le \alpha_\nu \le \beta_\nu \le ...,
\label{eq:interlacing}
\end{equation}
where $\alpha_1, \alpha_2,...$ are the zeros of one polynomial and
$\beta_1, \beta_2,...$ are the zeros of the other.
\end{itemize}
If, in the ordering of (\ref{eq:interlacing}), no equality sign occurs, then
$p(x)$ and $q(x)$ have {\em strictly interlacing zeros}.
\end{DEFI}
We use the following result of Hermite and Kakeya \cite{Rahman2002} to prove Theorem~\ref{theo:rige0}.
\begin{THEO}{\em (Hermite-Kakeya)}
Let $p(x)$ and $q(x)$ be non-constant polynomials in $x$ with real coefficients.
Then, $p(x)$ and $q(x)$ have strictly interlacing zeros if and only if, for
all $\mu$, $\lambda \in {\mathbb R}$ such that $\lambda^2 + \mu^2 > 0$, the polynomial
$g(x)=\lambda p(x) + \mu q(x)$ has simple, real zeros.
\end{THEO}
\noindent {\bf Proof of Theorem~\ref{theo:rige0}:}
We demonstrate in this proof that the numerator and denominator polynomials
in the RHS of (\ref{eq:partialdiff1}) have the same sign when evaluated at $x=r_i^2$.
Consider the three polynomials
$p_1(x)=D^{n-k+1}[G(x)]$, $p_2(x)=D^{n-k}\left\{G(x)/(x-S_j^2) \right\}$ and $q(x)=D^{n-k}[G(x)]$.
Note that $p_1(x)$ and $p_2(x)$ are of degree $(k-1)$, while $q(x)$ is of degree $k$.
We assert that
\begin{enumerate}
\item $p_1(x)$ and $q(x)$ have strictly interlacing zeros, and
\item $p_2(x)$ and $q(x)$ have strictly interlacing zeros.
\end{enumerate}
The first statement above is a straightforward consequence of the interlacing property of a polynomial with real zeros
and its derivative~\cite{Rahman2002}; we note that $p_1(x)=D[q(x)]$.
The second statement follows from a direct application of Hermite-Kakeya's Theorem to $p_2(x)$ and $q(x)$.
Consequently, for a given zero of $q(x)$, say $r_i^2$, there are an equal number of zeros of $p_1(x)$ and $p_2(x)$
to the left of $r_i^2$ on the real number line. Similarly,
there are an equal number of zeros of $p_1(x)$ and $p_2(x)$
to the right of $r_i^2$ on the real number line.
Lastly, note that the leading coefficients of all three polynomials (i.e., the coefficient of $x^k$ for $q(x)$
and the coefficients of $x^{k-1}$ for $p_1(x)$ and $p_2(x)$) are positive.
Therefore, $p_1(x)$ and $p_2(x)$ have the same sign when evaluated at $x=r_i^2$, and hence the RHS of (\ref{eq:partialdiff1})
is positive, completing the proof.
\qed
Finally, we present the main Theorem in this Section, which shows that the
choice of $S_j$'s that minimize the ratio $r_1^2/r_k^2$ is when the $S_j^2$'s are all equal.
\begin{THEO}
Under the assumptions of Theorem~\ref{theo:partialdiff}, the
ratio $r_1^2/r_k^2$ is minimized when $S_1^2= S_2^2 = ... = S_m^2$.
\label{theo:mincondition}
\end{THEO}
{\bf Proof:}
In order to minimize $r_1^2/r_k^2$, we consider its partial derivatives
with respect to the $S_j^2$'s. At the optimal location, we require the partial derivatives to be zero, i.e.,
\begin{eqnarray}
\frac{\partial}{\partial (S_j^2)} \left\{ \frac{r_1^2}{r_k^2} \right\} = 0
& \Rightarrow& \frac{1}{r_k^2} \left\{ r_k^2 \frac{\partial (r_1^2)}{\partial (S_j^2)} - r_1^2 \frac{\partial (r_k^2)}{\partial (S_j^2)} \right\} = 0, \nonumber \\
& \Rightarrow& \frac{r_1^2}{r_k^2} \left\{ \frac{1}{r_1^2} \frac{\partial (r_1^2)}{\partial (S_j^2)} - \frac{1}{r_k^2} \frac{\partial (r_k^2)}{\partial (S_j^2)} \right\} = 0, \nonumber \\
& \Rightarrow& \frac{1}{r_1^2} \frac{\partial (r_1^2)}{\partial (S_j^2)} = \frac{1}{r_k^2} \frac{\partial (r_k^2)}{\partial (S_j^2)}
\label{eq:mincondition}
\end{eqnarray}
for all the $S_j$'s.
We show that the above condition is satisfied with the choice $S_1^2=S_2^2=...=S_m^2$.
Assume that the $S_j^2$'s are all equal and non-zero, and denote their common value by $S^2$.
Because of symmetry, the quantities $\frac{\partial (r_i^2)}{\partial (S_j^2)}$ are independent of the choice of $j$, and we denote
the said quantity by $\frac{\partial (r_i^2)}{\partial (S^2)}$.
Consequently, the terms in the summation of LHS in the Euler homogeneity equation~(\ref{eq:eulerhomogeneous}) are all equal, and
hence
\begin{equation*}
\frac{\partial (r_i^2)}{\partial (S^2)} = \frac{r_i^2}{mS^2}.
\end{equation*}
Therefore, $\frac{1}{r_i^2} \frac{\partial (r_i^2)}{\partial (S^2)}$ is independent of the index $i$
and hence the condition (\ref{eq:mincondition}) is satisfied.
It remains to show that the optimal point just derived is a minimum.
A rigorous analysis to prove minimality involves the computation of the Hessian matrix~\cite{CalculusAlbert}
of $\frac{r_1^2}{r_k^2}$ with respect to the $S_j^2$'s. However, the analysis quickly turns intractable.
Instead, we fix $S_1^2=1$ and note that the only choice of the $S_j^2$'s for $j=2,3,...,k$ that satisfy the condition (\ref{eq:mincondition})
is when they are all equal to unity.
Thus, we can check the maxima or minima criteria by comparing the value of $\frac{r_1^2}{r_k^2}$ at $S_j^2=1$ to another
choice for the set $\{S_j^2\}$. We pick the set $S_1^2=1, S_2^2=0, S_3^2=0,...,S_m^2=0$ for the purpose of comparison, and
we immediately see that $\frac{r_1^2}{r_k^2} = \infty$ for this choice, because $r_k^2=0$. Therefore,
the optimal point we have determined is a minimum.
\qed
We remark that the above proof reveals that the ratio $r_{i_1}^2/r_{i_2}^2$ of any two roots $r_{i_1}^2$ and $r_{i_2}^2$
is minimized for a given $i_1$ and $i_2$ such that $i_1 < i_2$. Finally, we note that Theorem~\ref{theo:thompsonrip} together with Theorem~\ref{theo:mincondition} proves
Theorem~\ref{theo:main1}, which is the main result of this Section.
\subsection{Properties of the Structural Bound}
In this Section, we study the relationships between the structural bound given by Theorem~\ref{theo:main1}
and the parameters $n$, $m$, and $k$ of Compressed Sensing.
Clearly, the properties of the structural bound are tied to the properties of the
polynomial $f_k(x)$ as defined in Theorem~\ref{theo:main1}. We begin by
first expressing $f_k(x)$ in the standard polynomial form, by carrying out the
$(n-k)$'th order differentiation in (\ref{eq:fkx}). We obtain
\begin{equation}
f_k(x)=\sum_{j=0}^{k} (-1)^{k-j} \frac{{m \choose k-j} (n+j-k)!}{j!} x^j
\label{eq:fkxfull}
\end{equation}
and
\begin{equation}
g_k(x)=\sum_{j=0}^{k} (-1)^{k-j} \left( \sum S_1^2S_2^2...S_{k-j}^2 \right)
\frac{ (n+j-k)!}{j!} x^j.
\label{eq:gkxfull}
\end{equation}
We ask if the form of $f_k(x)$ is similar to one of classic polynomials in the literature~\cite{Rahman2002}.
The ratio of successive terms of the polynomial suggests that $f_k(x)$ is a Gauss Hypergeometric function of the
second kind:
\begin{eqnarray}
\frac{c_j}{c_{j+1}} & = & \frac{ (-1)^{k-j-1} {m \choose {k-j-1}} \frac{(n+j-k+1)!}
{(j+1)!} x^{j+1} } {(-1)^{k-j} {m \choose {k-j}} \frac{(n+j-k)!}
{j!} x^{j+1} } \nonumber \\
& = & \frac{(j-k)(j+n-k+1) } { (j+m-k+1) (j+1)} x. \nonumber
\end{eqnarray}
Therefore,
\begin{eqnarray}
f_k(x)& = &c \cdot _{2}F_{1}\left( -k, n-k+1; m-k+1; x \right), \nonumber
\end{eqnarray}
for some constant $c$.
There is very little known about the location of the zeros of
hyper-geometric functions of the kind described above~\cite{Ismail1991}.
However, algorithms to compute the zeros have been recently studied
\cite{Gil2004}.
\begin{figure*}
\centering
\subfigure[$k=4$] {\epsfysize = 60mm \epsffile{strucboundk4.eps} \label{fig-strucboundk4}}
\subfigure[$k=10$] {\epsfysize = 60mm \epsffile{strucboundk10.eps} \label{fig-strucboundk10}}
\caption{\sl Structural bound as a function of $n$. In each plot above, we fix $k$ to be a constant.
Each curve is obtained for a given value of $m$. }
\label{fig-strucbound}
\end{figure*}
To see the dependence of the structural bound on $n$, $m$ and $k$, we plot
$r_1^2/r_k^2$ as a function of $n$ in Figure~\ref{fig-strucbound}.
We assume that the $S_j^2$'s are all equal.
Several properties of the bound can be inferred from the plots. In particular,
\begin{enumerate}
\item For a given $m$ and $k$, the ratio $r_1^2/r_k^2$ increases when we increment $n$.
\item For a given $m$ and $k$, the ratio $r_1^2/r_k^2$ approaches a constant as we let
$n \longrightarrow \infty$. In other words, $r_1^2/r_k^2$ can be upper bounded by a constant
that is independent of $n$.
\item For a given $n$ and $m$, the ratio $r_1^2/r_k^2$ increases when we increment $k$.
\item For a given $n$ and $k$, the ratio $r_1^2/r_k^2$ decreases when we increment $m$.
\end{enumerate}
Each of the statements above is stated and proved in the form of Theorems below.
Alongside each Theorem, a result of the same flavor is proved for the RIP ratio for $\Phi$, if
such a result exists.
The following Theorem shows that when we keep $m$ and $k$ fixed,
the structural bound on the RIP ratio is an increasing function of $n$.
\begin{THEO}
Let $r_1^2 > r_2^2 > ... > r_k^2$ be the zeros of the polynomial $D^{n-k}[x^{n-m}(x-1)^m]$, and let
$t_1^2 > t_2^2 > ... > t_k^2$ be the zeros of the polynomial $D^{(n+1)-k}[x^{(n+1)-m}(x-1)^m]$.
Then,
\begin{equation*}
\frac{r_1^2}{r_k^2} \le \frac{t_1^2}{t_k^2}.
\end{equation*}
\label{theo:incrn}
\end{THEO}
{\bf Proof:}
We have
\begin{eqnarray}
D^{(n+1)-k}[x^{(n+1)-m}(x-1)^m] &=& D^{n-k}[D[x^{(n+1)-m}(x-1)^m]] \nonumber \\
&=& D^{n-k}[ (n+1-m)x^{n-m}(x-1)^{m} + m x^{(n+1)-m}(x-1)^{m-1} ] \nonumber \\
&=& D^{n-k}\left[x^{n-m}(x-1)^{m-1} \left\{ (n+1-m)(x-1) + mx \right\} \right] \nonumber \\
&=& (n+1) D^{n-k}\left[x^{n-m}(x-1)^{m-1} \left( x- \left( 1-\frac{m}{n+1} \right) \right) \right].
\label{eq:ntonp1}
\end{eqnarray}
Comparing (\ref{eq:ntonp1}) and (\ref{eq:fkx}) reveals that the polynomial $f_k(x)$ evaluated for $n+1$ and $m$ is of the same form (up to a constant)
of $g_k(x)$ evaluated for $n$ and $k$ with all singular values equal except one with value $S_m^2=\left( 1-\frac{m}{n+1} \right)$.
Because the singular values are not all equal, Theorem~\ref{theo:mincondition} ensures that $\frac{r_1^2}{r_k^2} \le \frac{t_1^2}{t_k^2}$.
This completes the proof of Theorem~\ref{theo:incrn}.
\qed
\begin{THEO}
Let $\Phi_2$ be an $m \times (n+1)$ sized matrix over ${\mathbb R}$ or ${\mathbb C}$, and let
$\Phi_1$ be an $m \times n$ submatrix of $\Phi_2$.
Then, the RIP ratios of the two matrices satisfy
$R(\Phi_1,k) \le R(\Phi_2,k)$.
\label{theo:incrnphi}
\end{THEO}
{\bf Proof:} Let $\Lambda_1$ be the set of all squared singular values $\{s_{p,i}^2\}$ of all submatrices of
$\Phi_1$, and let $\Lambda_2$ be the set of all squared singular values of all submatrices of
$\Phi_2$.
Since every $m \times k$ submatrix of $\Phi_1$ is also a submatrix of $\Phi_2$, we have
$\Lambda_1 \subset \Lambda2$. Therefore, $\max\{ \Lambda_1\} \le \max\{ \Lambda_2 \}$ and
$\min\{ \Lambda_1\} \ge \min\{ \Lambda_2 \}$. Because $R(\Phi_1,k) = \max\{ \Lambda_1\} / \min\{ \Lambda_1\}$
and $R(\Phi_2,k) = \max\{ \Lambda_2\} / \min\{ \Lambda_2\}$, the result of Theorem~\ref{theo:incrnphi}
follows as a consequence.
\qed
The following Theorem shows that the structural bound can itself be bounded by a quantity that is independent of $n$.
\begin{THEO}
Let $r_1^2 > r_2^2 > ... > r_k^2$ be the zeros of the polynomial $D^{n-k}[x^{n-m}(x-1)^m]$. Then,
\begin{equation}
\frac{r_1^2}{r_k^2} \le \frac{(m k)^k (m-k)!}{m!}.
\label{eq:boundratior1rk}
\end{equation}
\label{theo:boundratior1rk}
\end{THEO}
{\bf Proof:}
We establish the result by deriving an upper bound for $r_1^2$ and a lower bound for $r_k^2$.
Applying Vi\'ete's Theorem to the polynomial in (\ref{eq:fkxfull}), we have
\begin{equation*}
r_1^2 + r_2^2 + ... + r_k^2 = \frac{mk}{n}.
\end{equation*}
Since each term in the LHS of the above equation is positive, we have
\begin{equation}
r_1^2 \le \frac{mk}{n},
\label{eq:boundr1}
\end{equation}
giving us an upper bound on $r_1^2$.
To derive a lower bound on $r_k^2$, we begin by using Vi\'ete's Theorem for the constant term of (\ref{eq:fkxfull}),
involving the product $r_1^2 r_2^2 ... r_k^2$:
\begin{equation*}
r_1^2 r_2^2 ... r_k^2 = \frac{{m \choose k}} {{n \choose k}}.
\end{equation*}
Since $r_i^2 > r_k^2$ for all $i <k$, we have the inequality
\begin{equation*}
r_k^2 (r_1^2)^{k-1} \ge \frac{{m \choose k}} {{n \choose k}}.
\end{equation*}
Applying (\ref{eq:boundr1}) to the above inequality, we have
\begin{equation*}
r_k^2 \left( \frac{mk}{n}\right)^{k-1} \ge r_k^2 (r_1^2)^{k-1} \ge \frac{{m \choose k}} {{n \choose k}}.
\end{equation*}
Therefore, we have
\begin{eqnarray}
r_k^2 & \ge & \frac{{m \choose k}} {{n \choose k}} \frac{n^{k-1}}{(mk)^{k-1}} \nonumber \\
& = & \frac{m! (n-k)! n^{k-1}} {(m-k)! n! (mk)^{k-1}} \nonumber \\
& = & \frac{m!} {n (m-k)! (mk)^{k-1}} \frac{(n-k)! n^k}{n!} \nonumber \\
& \ge & \frac{m!} {n (m-k)! (mk)^{k-1}}, \label{eq:boundrk}
\end{eqnarray}
where the last inequality (\ref{eq:boundrk}) is a consequence of $n^k \ge n! /(n-k)!$.
Combining the inequalities (\ref{eq:boundr1}) and (\ref{eq:boundrk}) by taking the ratio (noting that both
inequalities have positive LHS and RHS), we obtain the inequality in (\ref{eq:boundratior1rk}),
completing the proof of Theorem~\ref{theo:boundratior1rk}.
\qed
Although Theorem~\ref{theo:boundratior1rk} gives an upper bound for the structural bound $r_1^2/r_k^2$ that is independent
of $n$, we cannot bound the RIP ratio $R(\Phi, k)$ as we increase $n$.
This is because $R(\Phi,k)$ necessarily increases with $n$. Consequently, as we increase $n$, the structural bound
becomes more loose (this can be seen in Figure~\ref{fig-bounds-results}).
The packing bound, discussed below in Section~\ref{sec:packing}, provides a tighter bound for
large $n$ because it captures the growth of $R(\Phi,k)$ with increasing $n$.
\begin{THEO}
Let $r_1^2 > r_2^2 > ... > r_k^2$ be the zeros of the polynomial $D^{n-k}[x^{n-m}(x-1)^m]$, and let
$t_1^2 > t_2^2 > ... > t_k^2$ be the zeros of the polynomial $D^{n-k}[x^{n-m+1}(x-1)^{m-1}]$.
Then,
\begin{equation*}
\frac{r_1^2}{r_k^2} \le \frac{t_1^2}{t_k^2}.
\end{equation*}
\label{theo:incrm}
\end{THEO}
{\bf Proof:}
Reducing $m$ by one is equivalent to the following operation: set $S_m^2=0$.
The proof of the above Theorem follows therefore
from a straightforward application of Theorem~\ref{theo:mincondition}.
\qed
\begin{THEO}
Let $\Phi_2$ be an $(m+1) \times n$ sized matrix over ${\mathbb R}$ or ${\mathbb C}$, and let
$\Phi_1$ be an $m \times n$ submatrix of $\Phi_2$.
Then, the RIP ratios of the two matrices satisfy
$R(\Phi_1,k) \le R(\Phi_2,k)$.
\label{theo:incrmphi}
\end{THEO}
{\bf Proof:}
Let $p \in \{ 1,2,...,{n \choose k}\}$ be the index that identifies the set of $k$ columns
that are selected from the complete matrix to form a submatrix with $k$ columns.
Then, $(\Phi_1)_p$ is a submatrix of $(\Phi_2)_p$, and consequently,
the maximum (minimum) singular value of $(\Phi_1)_p$ is smaller (greater) than the
maximum (minimum) singular value of $(\Phi_2)_p$ by the interlacing Theorem for matrices~\cite{RCThompson9}.
Thus Theorem~\ref{theo:incrmphi} is established.
\qed
\begin{THEO}
Let $r_1^2 > r_2^2 > ... > r_k^2$ be the zeros of the polynomial $D^{n-k}[x^{n-m}(x-1)^m]$, and let
$t_1^2 > t_2^2 > ... > t_{k+1}^2$ be the zeros of the polynomial $D^{n-k-1}[x^{n-m+1}(x-1)^{m-1}]$.
Then,
\begin{equation*}
\frac{r_1^2}{r_k^2} \le \frac{t_1^2}{t_{k+1}^2}.
\end{equation*}
\label{theo:incrk}
\end{THEO}
{\bf Proof:}
Since $D^{n-k-1}[x^{n-m+1}(x-1)^{m-1}] = D\left[ D^{n-k}[x^{n-m}(x-1)^m] \right]$, the roots of
the two polynomials weakly interlace, as a consequence of the interlacing theorem for a polynomial and
its derivative~\cite{Rahman2002}. Therefore, $r_1^2 \le t_1^2$ and $r_k^2 \ge t_{k+1}^2$ and the proof
of Theorem~\ref{theo:incrk} follows as a consequence.
\qed
Finally, we state a similar Theorem for matrices.
\begin{THEO}
Let $\Phi$ be an $m \times n$ matrix over ${\mathbb R}$ or ${\mathbb C}$.
Then, the RIP ratios satisfy
$R(\Phi,k) \le R(\Phi,k+1)$.
\label{theo:incrkphi}
\end{THEO}
The proof of Theorem~\ref{theo:incrkphi} is identical to the proof of Theorem~\ref{theo:incrmphi}.
\subsection{Geometric Interpretation}
Recall the geometric interpretation of the SVD: The matrix $\Phi$ with SVD $\Phi = U S V^T$ can be represented as a hyperellipse
of dimension $m$ embedded in ${\mathbb R}^n$. The axes of the hyperellipse are aligned with the column vectors of $V$ with the length of each semi-axis
equal to the corresponding singular value. Denote this hyperellipse by $E(\Phi)$. It is well known~\cite{KalmanSVD} that $\| \Phi x\|_2$ is equal
to the magnitude of the projection of the vector $x$ onto the hyperellipse $E(\Phi)$. Furthermore, the column vectors of $U$ describe the
orientation of the hyperellipse in ${\mathbb R}^m$, which is the image of $\Phi x$ of the unit sphere in ${\mathbb R}^n$.
That is, for $\|x\|_2=1$ we have
\begin{equation}
{n-q \choose k-q} \sum_{q-\mbox{wise}}(S_1 S_2 ... S_q)^2 = \sum_p \left\{ \sum_{q-\mbox{wise}} (s_{p,1} s_{p,2}...s_{p,q})^2 \right\},
\label{eq:GGPT1}
\end{equation}
where the $q$-wise summation in (\ref{eq:GGPT1}) is the sum of all the terms that are obtained by multiplying $q$ unique
singular values.\footnote{For example, consider $q=2$, $m=4$ and $k=3$ for illustration. Then,
$\sum_{q-\mbox{wise}}(S_1 S_2 ... S_q)^2 = (S_1 S_2)^2 + (S_1 S_3)^2 + (S_1 S_4)^2 + (S_2 S_3)^2 + (S_2 S_4)^2 + (S_3 S_4)^2$, and $\sum_{q-\mbox{wise}} (s_{p,1} s_{p,2}...s_{p,q})^2 = (s_{p,1} s_{p,2})^2 + (s_{p,1} s_{p,3})^2 + (s_{p,2} s_{p,3})^2$.}
Equation~(\ref{eq:GGPT1}) relates the dimensions of $E(\Phi)$ to the dimensions of
the collection of $k$-dimensional hyperellipses $E(\Phi_p)$ corresponding to each $\Phi_p$.
Note that each of the hyperellipses $E(\Phi_p)$ lie in a $k$-dimensional subspace
spanned by $k$ canonical basis vectors.
A particularly interesting case is when
$m=k=q$, for which (\ref{eq:GGPT1}) reduces to
\begin{equation}
(S_1 S_2 ... S_k)^2 = \sum_p (s_{p,1} s_{p,2}...s_{p,k})^2.
\label{eq:gpt}
\end{equation}
The above result for $m=k=q$ is equivalent to the Generalized Pythagorean Theorem
(GPT)~\cite{GPT1,GPT2}. The GPT states that
the square of the $k$-volume of a $k$-dimensional parallelepiped embedded in an $n$-dimensional Euclidean space is equal to the
sum of the squares of the $k$-volumes of the ${n \choose k}$ projections of the parallelepiped on to the distinct $k$-dimensional
subspaces spanned by the canonical basis vectors.
Equation~(\ref{eq:gpt}) implies that the statement of the GPT can be directly carried over from parallelepipeds to hyperellipses.
Equation~(\ref{eq:GGPT1}) extends GPT to arbitrary values of $m$, $k$ and $q$.\footnote{Note
that the hyperellipses $E(\Phi_p)$ are the projections of $E(\Phi)$ onto the canonical $k$-subspaces
only when $m=k$.}
The form of the equation motivates us to define the $q$-volume of an ellipse of intrinsic dimension that is greater than $q$ as follows.
\begin{DEFI}
Consider a hyperellipse $H$ of dimension $d>q$ with semi-axes $a_1, a_2, ..., a_d$.
The $q$-volume of $H$, denoted by $\mbox{Vol}_q(H)$, is defined as
\begin{equation*}
\mbox{Vol}_q(H) \triangleq \sqrt{ \sum_{q-\mbox{wise}} (a_1 a_2 ... a_q)^2 }.
\end{equation*}
\end{DEFI}
Equation~(\ref{eq:GGPT1}) therefore relates the $q$-volumes of $E(\Phi)$ to the $q$-volumes of $E(\Phi_p)$ in the following manner: the
square of the
$q$-volume of $E(\Phi)$ is proportional to the sum of the squares of the $q$-volumes of $E(\Phi_p)$.
\subsection{Structural Bound for $k=2$}
In this subsection, we study the structural bound for the specific case of $k=2$.
The motivation for studying this case are many fold. First, $k=2$ is the smallest
non-trivial case to investigate the RIP ratio. For the case $k=1$,
any matrix $\Phi$ that has equi-normed columns satisfies $R(\Phi,1)=1$, and therefore the structural
bound is trivially $1$ for any $n$ and $m$.
Secondly, the roots of the polynomial~(\ref{eq:fkxfull}) can be explicitly evaluated, providing
an avenue for analysis.
Lastly and most importantly, designing good CS matrices for $k=2$ can be shown to be equivalent to well-known problems
in coding theory.
For $k=2$, the form of (\ref{eq:fkxfull}) and (\ref{eq:gkxfull}) reduce to
\begin{equation}
\frac{n(n-1)}{2}x^2 - \left( \sum_{j=1}^{m} S_j^2 \right) (n-1)x +
\left( \sum_{1 \le j_1 < j_2 \le m} S_{j_1}^2S_{j_2}^2 \right) = 0
\label{eq:quadratic_diffS}
\end{equation}
and
\begin{equation}
\frac{n(n-1)}{2}x^2 - (n-1)mx + \frac{m(m-1)}{2} = 0.
\label{eq:f2x}
\end{equation}
We focus on (\ref{eq:f2x}), because we are interested in universal bounds for the RIP ratio.
The roots of (\ref{eq:f2x}) can be computed as
\begin{equation*}
r_1^2, r_2^2 = \frac{m}{n} \pm \frac{1}{n}\sqrt{\frac{m(n-m)}{n-1}},
\end{equation*}
and the structural bound is thus
\begin{equation*}
\frac{r_1^2}{r_2^2} = \frac{1 + \sqrt{\frac{n-m}{m(n-1)}} } {1 - \sqrt{\frac{n-m}{m(n-1)}}}.
\end{equation*}
Note that as $n \longrightarrow \infty$, the above equation reduces to
\begin{eqnarray}
\lim_{n \longrightarrow \infty}\frac{r_1^2}{r_2^2} \mbox{~~}= \mbox{~~} \frac{ 1 + \sqrt{ \frac{1}{m}} } {1 - \sqrt{
\frac{1}{m}} }
\mbox{~~}=\mbox{~~} \frac{\left( 1 + \sqrt{\frac{1}{m}} \right)^2}{1-\frac{1}{m}}.
\label{eq:limninftyk2}
\end{eqnarray}
Recall that we derived an upper bound (\ref{eq:boundratior1rk}) on $\frac{r_1^2}{r_2^2}$ that is applicable
for any $m$, $k$ and as $n \longrightarrow \infty$. Substituting $k=2$ in (\ref{eq:boundratior1rk}), we obtain
\begin{eqnarray}
\frac{r_1^2}{r_k^2} \mbox{~~} \le \mbox{~~} \frac{(2m)^2(m-2)!}{m!}
\mbox{~~} = \mbox{~~} \frac{4m^2}{m(m-1)}
\mbox{~~} = \mbox{~~} \frac{4}{1-\frac{1}{m}}.
\label{eq:universalk2}
\end{eqnarray}
Comparison of (\ref{eq:limninftyk2}) and (\ref{eq:universalk2}) reveals that (\ref{eq:limninftyk2})
offers a tighter bound on $r_1^2 / r_k^2$ than (\ref{eq:universalk2}).
We now make some interesting
connections between good CS matrices for $k=2$ and coding theory. The key result that
provides the segue is Theorem~\ref{theo:everymequalcols}.
\begin{THEO}
Let $A=\left[a_1 ~~ a_2 ~~ a_3 ~ ... ~a_n \right]$ be an $m \times n$ matrix over ${\mathbb R}$ or ${\mathbb C}$ comprising
$n \ge 2$ columns $a_1 , a_2, ... a_n$
of size $m \times 1$, with $\| a_i \|_2 >0$.
Construct the $m \times n$ matrix
$B$
as
\begin{equation}
B = \left[\frac{a_1}{\|a_1\|_2} ~~ \frac{a_2}{ \|a_2\|_2} ~~ \frac{a_3}{ \|a_3\|_2} ~ ... ~ ~~ \frac{a_n}{ \|a_n\|_2}\right],
\label{eq:bfroma}
\end{equation}
obtained by
scaling every column of $A$ independently so that the $\ell_2$ norm of every column of $B$ is unity.
Then, the RIP ratios of $A$ and $B$ satisfy $R(B,2) \le R(A,2)$.
\label{theo:everymequalcols}
\end{THEO}
{\bf Proof:}
We prove the Theorem by considering two cases: $n=2$ and $n >2$. \\
{\bf Case 1: $n=2$.}\\
Since $A$ has only two columns, the only submatrix of $A$ of size $m \times 2$ is $A$ itself. Therefore,
the RIP ratio is simply the ratio of the square of the two singular values of $A$.
Let the $\ell_2$ norms of column vectors $a_1$ and $a_2$ be $\|a_1\|_2 = d_1$ and $\|a_2\|_2 = d_2$.
Let the angular distance between $a_1$ and $a_2$ be $\theta$, given by $\cos(\theta) = | \left<a_1 \cdot a_2 \right> | /(d_1 d_2)$,
where $| \left<a_1 \cdot a_2 \right> |$ is the absolute value of the dot product of $a_1$ and $a_2$.
Let $S_1^2$ and $S_2^2$ be the squared singular values of $A$, with $S_1^2 \ge S_2^2$.
Since the squared singular values of $A$ are the eigen values of the grammian matrix $A^H A$, we compute $A^H A$
as
\begin{equation}
A^H A =
\left( \begin{array}{cc}
d_1^2 & d_1 d_2 \cos (\theta) \\
d_1 d_2 \cos (\theta) & d_2^2 \end{array} \right).
\label{eq:gramA}
\end{equation}
Thus $S_1^2$ and $S_2^2$ are the zeros of the characteristic polynomial of
(\ref{eq:gramA}) given by $x^2 -(d_1^2 + d_2^2)^2 + \left( d_1 d_2 \right)^2 \sin^2 (\theta)$.
Computing the roots of this polynomial yields
\begin{equation*}
S_1^2= \frac{1}{2} \left[ d_1^2+d_2^2 + \sqrt{\left( d_1^2+d_2^2 \right)^2 - 4 \left(d_1 d_2 \right)^2 \sin^2 (\theta)} \right]
\end{equation*}
and
\begin{equation*}
S_2^2= \frac{1}{2} \left[ d_1^2+d_2^2 - \sqrt{\left( d_1^2+d_2^2 \right)^2 - 4 \left(d_1 d_2 \right)^2 \sin^2 (\theta)} \right].
\end{equation*}
The RIP ratio of $A$ is given by
\begin{eqnarray}
R(A,2) = \frac{S_1^2}{S_2^2} &=& \frac{ d_1^2+d_2^2 + \sqrt{\left( d_1^2+d_2^2 \right)^2 - 4 \left(d_1 d_2 \right)^2 \sin^2 (\theta)} }{d_1^2+d_2^2 - \sqrt{\left( d_1^2+d_2^2 \right)^2 - 4 \left(d_1 d_2 \right)^2 \sin^2 (\theta)} } \nonumber \\
& = & \frac{1 + \sqrt{1 - \nu \sin^2 (\theta)} }{1 - \sqrt{1 - \nu \sin^2 (\theta)}}, \label{eq:apgp}
\end{eqnarray}
where
\begin{equation*}
\nu = \frac{ \left( 2d_1 d_2 \right) }{d_1^2 + d_2^2}
\end{equation*}
is the ratio of the geometric mean and the arithmetic mean of $d_1^2$ and $d_2^2$.
From (\ref{eq:apgp}), we infer that in order to minimize $R(A,2)$ for a fixed $\theta$, we require $\nu$ to be as large
as possible. Since $\nu$ is the ratio of geometric mean and arithmetic mean of $d_1^2$ and $d_2^2$,
maximum value of $\nu$ is attained when $d_1 = d_2$, yielding $\nu_{\max} = 1$.
Since the angular separation of the column vectors remains invariant while constructing $B$ from $A$ using
(\ref{eq:bfroma}), and we have equal column norms in $B$, we infer that $R(B,2) \le R(A,2)$ from the above arguments.
Thus we have proved Theorem~\ref{theo:everymequalcols} for the case $n=2$.
Note that when $d_1=d_2 =d$, we have
\begin{equation}
S_1^2 = d^2 \left(1 + \cos (\theta) \right),
\label{eq:optimals1}
\end{equation}
\begin{equation}
S_2^2 = d^2 \left(1 - \cos (\theta) \right),
\label{eq:optimals2}
\end{equation}
and
\begin{equation}
\frac{S_1^2}{S_2^2} = \frac{1 + \cos(\theta)}{1 - \cos(\theta)} = \cot^2 \frac{\theta}{2}.
\label{eq:optimalRd1d2}
\end{equation}
While Case 1 is applicable to matrices with two columns, we extend the result
to include any $m \times n$ matrix in Case 2. \\
{\bf Case 2: $n > 2$.} \\
Consider the following three $m \times 2$ sized submatrices $A_{p_0}$, $A_{p_1}$, and $A_{p_2}$ of $A$, where
$p_0, p_1, p_2$ are indices from $1,2,...,{n \choose 2}$:
\begin{enumerate}
\item $A_{p_0}$ is obtained by selecting the two columns of $A$ that have the minimum angular distance among all
pairs of column vectors of $A$. Let the singular values of $A_{p_0}$ be $s_{p_0,1}$ and $s_{p_0,2}$
with $s_{p_0,1} \ge s_{p_0,2}$.
\item $A_{p_1}$ is the $m \times 2$ submatrix of $A$ whose largest singular value $s_{p_1,1}$ is the maximum of all
singular values of submatrices of $A$ of size $m \times 2$. In other words,
\begin{equation*}
p_1 = \arg \max_{p} \{s_{p,1}\}, \mbox{~~~~~~~~$p = 1,2,...,{n \choose 2}$}.
\end{equation*}
Let the singular values of $A_{p_1}$ be $s_{p_1,1}$ and $s_{p_1,2}$ with $s_{p_1,1} \ge s_{p_1,2}$.
\item $A_{p_2}$ is the $m \times 2$ submatrix of $A$ whose smallest singular value $s_{p_2,2}$ is the minimum of all
singular values of submatrices of $A$ of size $m \times 2$. In other words,
\begin{equation*}
p_2 = \arg \min_{p} \{s_{p,2}\}, \mbox{~~~~~~~~$p = 1,2,...,{n \choose 2}$}.
\end{equation*}
Let the singular values of $A_{p_2}$ be $s_{p_2,1}$ and $s_{p_2,2}$ with $s_{p_2,1} \ge s_{p_2,2}$.
\end{enumerate}
We also consider an $m \times 2$ submatrix $B_{p_0}$ of $B$, using the same index $p_0$ defined above.
The angular separation between column vectors are invariant to scaling of columns;
therefore $B_{p_0}$ contains the
two columns of $B$ that have the minimum angular distance among all pairs of column vectors of $B$.
We denote the singular values of $B_{p_0}$ as $t_1$ and $t_2$ with $t_1 \ge t_2$. \\
From the definition of RIP ratio, we have
\begin{equation}
R(A,2) = \frac{s_{p_1,1}^2}{s_{p_2,2}^2}.
\label{eq:ripratioa}
\end{equation}
Since $s_{p_1,1} \ge s_{p_0,1}$ and $s_{p_2,2} \le s_{p_0,2}$, we have
\begin{equation}
\frac{s_{p_1,1}^2}{s_{p_2,2}^2} \ge \frac{s_{p_0,1}^2}{s_{p_0,2}^2}.
\label{eq:sp0sp1sp2}
\end{equation}
Invoking the result we have established from Case 1 for the $m \times 2$ sized
matrices $A_{p_0}$ and $B_{p_0}$, we have $R(A_{p_0},2) \ge R(B_{p_0},2)$ and so
\begin{equation}
\frac{s_{p_0,1}^2}{s_{p_0,2}^2} \ge \frac{t_1^2}{t_2^2}.
\label{eq:absubmatrix}
\end{equation}
Let us consider the RIP ratio of the matrix $B$. Since every column of $B$ has equal norm, we assert that the
RIP ratio of $B$ is governed completely by the submatrix $B_{p_0}$. To see this, we first note that when the column norms are equal,
the condition number of an $m \times 2$ matrix
is dictated only by the angular separation of its two constituent column vectors. Specifically,
if the two columns of an $m \times 2$ matrix have equal $\ell_2$ norm of $d$ and their angular
separation is $\theta$, then the squared singular values are given by (\ref{eq:optimals1}) and (\ref{eq:optimals2}).
Furthermore, the squared condition number is given by (\ref{eq:optimalRd1d2}). The squared
condition number is a monotonically decreasing function of $\theta$ in the range
$0 \le \theta \le \pi/2$.
Therefore, the RIP ratio of $B$ is given by
\begin{equation}
R(B,2) = \frac{t_1^2}{t_2^2}.
\label{eq:ripratiob}
\end{equation}
Using the results of (\ref{eq:ripratioa}), (\ref{eq:sp0sp1sp2}), (\ref{eq:absubmatrix}) and (\ref{eq:ripratiob}) and cascading
the inequalities, we obtain
$R(A,2) \ge R(B,2)$, and so the proof of Theorem~\ref{theo:everymequalcols} is complete.
\qed
The above Theorem has important consequences for the RIP ratio of order $k=2$.
\begin{REMA}
Theorem~\ref{theo:everymequalcols} reveals that among the set of all matrices that can be obtained by scaling each column
of a given matrix independently, the RIP ratio for $k=2$ is minimized when the column norms are all equal.
\end{REMA}
We use the above results to study the properties of a CS matrix $\Phi$ that attains the structural bound for $k=2$.
First, Theorem~\ref{theo:main1} (third statement) requires that $\Phi_p$ has the same pair of squared singular values for
all $p$. Second, we deduce from Theorem~\ref{theo:everymequalcols} that the columns of $\Phi$ are all equi-normed. If the contrary were true, then
equalizing the column norms will yield a matrix with smaller RIP ratio, which violates Theorem~\ref{theo:main1}. Since this result is of importance,
we state it in the form of a Theorem.
\begin{THEO}
If an $m \times n$ matrix $\Phi$ over ${\mathbb R}$ or ${\mathbb C}$
satisfies (\ref{eq:RIPbound1}) with equality, then $\Phi$ has equi-normed columns.
\label{theo:equinormed}
\end{THEO}
Furthermore, we show that $\Phi$ that satisfies (\ref{eq:RIPbound1}) with equality is an equi-angular tight frame (ETF)~\cite{TroppEquiangular}.
From \cite{TroppEquiangular,WelchEquiangular}, we list the three conditions for a matrix $A$ to be an ETF:
\begin{enumerate}
\item The columns of $A$ are unit normed,
\item The absolute values of the dot product of every pair of columns of $A$ are same, i.e., the columns are equi-angular.
\item $A A^*=(n/m)I_{m \times m}$.
\end{enumerate}
\begin{THEO}
Let an $m \times n$ matrix $\Phi$ over ${\mathbb R}$ or ${\mathbb C}$
satisfy (\ref{eq:RIPbound1}) with equality, and let $\Phi$ be scaled
such that its $m$ squared singular values are
equal to $n/m$ each, i.e., $S_1^2=S_2^2=...S_m^2=n/m$. Then, $\Phi$ is an ETF.
\label{theo:ripetf}
\end{THEO}
{\bf Proof:}
Let $\Phi=U S V^*$ be the singular value decomposition for $\Phi$.
We check the three conditions for ETF, starting with the third condition.
We have $A A^* = U S V^* V S U^* = U S^2 U^*=(n/m)UIU^*=(n/m)I$, satisfying the third condition.
The second condition is satisfied because every pair of columns have the same set of
singular values. If the angle between two equi-normed columns is $\theta$ and the common norm is
$d$, then the squared singular values of the $m \times 2$ sized submatrix comprising only of the two said columns
are given by (\ref{eq:optimals1}) and (\ref{eq:optimals2}).
To verify the first condition, we note that the norms of each column of $\Phi$ are the same, say $d^2$. It remains to
show that this norm is unity.
Substituting $S_1^2=S_2^2=...S_m^2=n/m$ in (\ref{eq:quadratic_diffS}),
we see that $s_1^2$ and $s_2^2$ are the roots of the quadratic equation
\begin{equation}
x^2 - 2x + \frac{n(m-1)}{m(n-1)} = 0.
\label{eq:welchfoundation}
\end{equation}
Therefore the sum of the roots of the quadratic equation is given by
\begin{equation}
s_1^2 + s_2^2=2.
\label{eq:oned}
\end{equation}
From (\ref{eq:optimals1}) and (\ref{eq:optimals2}), we have
\begin{equation}
s_1^2+s_2^2=2d^2,
\label{eq:twod}
\end{equation}
and hence we infer that
$d^2=1$ by comparing
(\ref{eq:oned}) and (\ref{eq:twod}). Thus $\Phi$ satisfies all three
conditions for an ETF.
\qed
The relationship between the structural bound and ETFs can be used to make statements about the
set of allowable pairs $(m,n)$ that meet the structural bound. Results from \cite{TroppEquiangular}
reveal that an ETF of size $m \times n$ exists only when $n \le \frac{1}{2} m (m+1)$ for real ETFs and
$n \le m^2$ for complex ETFs. In addition, $n$ and $m$ should satisfy strict integer constraints.
\subsection{Structural Bound as Extension of the Welch Bound}
\begin{DEFI}
The {\em coherence} of a matrix $A$, denoted by $\mu(A)$ is defined as the largest absolute inner product between any two columns $a_i$, $a_j$ of $A$. That is, for $A$ with $n$ columns,
\begin{equation}
\mu(A) = \max_{i,j \in \{1,2,...n\}, i \ne j} \frac{| \left< a_i \cdot a_j \right>|}{\|a_i\|_2 \|a_j\|_2}.
\label{eq:welchdef}
\end{equation}
\end{DEFI}
Recall the classical result of Welch that states
that the coherence of a $m \times n$ matrix in ${\mathbb R}$ or ${\mathbb C}$ is always greater than or equal to
$\sqrt{\frac{n-m}{m(n-1)}}$. It is well known that $\Phi$ is an ETF if and only if the
coherence of $\Phi$ satisfies the Welch bound with equality~\cite{WelchEquiangular,TroppEquiangular}. As a consequence of
Theorem~\ref{theo:ripetf}, we infer that if $\Phi$
satisfies (\ref{eq:RIPbound1}) with equality, then the coherence of $\Phi$ meets the Welch bound.
We make the claim that the structural bound for $k>2$ extends the Welch bound from pairs of column
vectors of a matrix to $k$-tuples of column vectors. We use the
least singular value of a submatrix as a way to extend the notion of coherence to a
$k$-tuple of vectors. Specifically, we wish to maximize the (normalized) least singular value of the submatrices
in order to keep the submatrices as far apart as possible, in some sense.\footnote{Without normalizing the least
singular value, it can be increased arbitrarily by simply scaling the matrix. As we shall see, the normalization is done
based on the largest singular value of $\Phi$.}
However, Theorem~\ref{theo:thompsonrip}
asserts that the least singular value has an upper bound. Based on this insight, we extend the Welch bound explicitly in the following Theorem.
\begin{THEO} {\em (Extension of Welch Bound)}
Let $\Phi$ be an $m \times n$ matrix over ${\mathbb R}$ or ${\mathbb C}$ with $0 < m < n$. Let $k$
be an integer such that $0 < k < m$, and let $f_k(x)$ be the $k$'th degree polynomial given by
$f_k(x) = D^{n-k} \left[ x^{n-m}(x-1)^m \right]$.
Let $r_1^2 \ge r_2^2 \ge ... \ge r_k^2$ be the zeros of $f_k(x)$.
Let $S_1$ be the largest singular value of
$\Phi$. Let $s_{p,1} \ge s_{p,2} \ge... \ge s_{p,k},$ be the $k$ singular values of the $m \times k$ sized submatrix $\Phi_p$ of $\Phi$, where $p \in \{1,2,...,{n \choose k}\}$ is the index that identifies the submatrix.
Then the smallest $s_{p,k}$, after normalization by $S_1^2$, is bounded by
\begin{equation}
\left( \frac{\min_{p} \{ s_{p,k}^2 \}}{S_1^2} \right) \le r_k^2.
\label{eq:welchextensionresult}
\end{equation}
\label{th:welchextension}
\end{THEO}
{\bf Proof:}
Let $S_1 \ge S_2 \ge ... \ge S_m \ge 0$ be the $m$ singular values of
$\Phi$.
Define two more polynomials $g_k(x)$ and $h_k(x)$, both of degree $k$, given by
\begin{equation*}
g_k(x) = D^{n-k} \left[ x^{n-m} (x-S_1^2)(x-S_2^2)...(x-S_m^2) \right] ~~\mbox{ and }~~
h_k(x) = D^{n-k} \left[ x^{n-m} (x-S_1^2)^m \right].
\end{equation*}
Let the zeros of $g_k(x)$ be $t_1^2 \ge t_2^2 \ge ... \ge t_k^2$, and the zeros of
$h_k(x)$ be $u_1^2 \ge u_2^2 \ge ... \ge u_k^2$. \\
Based on the result of Theorem~\ref{theo:thompsonrip}, we have
\begin{equation}
\min_{p} \{ s_{p,k}^2 \} \le t_k^2.
\label{eq:proofwelchextensioneq1}
\end{equation}
From Theorem~\ref{theo:rige0}, we see that $t_i^2 \le u_i^2$ for all $i=1,2,...,k$, because we can think of the polynomial $h_k(x)$ as being obtained by increasing each of $S_2^2, S_3^2,...,S_m^2$ to $S_1^2$. Theorem~\ref{theo:rige0} guarantees
that the zeros of $h_k(x)$ are greater than the corresponding zeros of $g_k(x)$.
Therefore, $t_k^2 \le u_k^2$ and so
\begin{equation}
\min_{p} \{ s_{p,k}^2 \} \le u_k^2.
\label{eq:proofwelchextensioneq2}
\end{equation}
While (\ref{eq:proofwelchextensioneq1}) is a tighter bound than (\ref{eq:proofwelchextensioneq2}), the latter equation
has the advantage that it depends only on $S_1^2$ and holds for all values of $S_i^2 \le S_1^2$, for $i=2,3,...,m$.
In the final step of the proof, we use the result $u_i^2 = S_1^2 r_i^2$ which follows
by noting that $h_k(x)$ is obtained by making the change of variable $x \rightarrow S_1 x$ in $f_k(x)$. Substituting in (\ref{eq:proofwelchextensioneq2}),
we obtain $\min_{p} \{ s_{p,k}^2 \} \le S_1^2 r_k^2$ and the proof is complete.
\qed
Based on the proof of Theorem~\ref{theo:thompsonrip}, we infer that (\ref{eq:welchextensionresult}) holds with
equality if and only if $s_{p,k} = S_1^2 r_k^2$ for all $p \in \{1,2,...,{n \choose k}\}$. This extends the notion of
ETFs, which for $k=2$, require only the
angular separation between every pair of column vectors to be the same.
Note that a necessary condition for the equality of (\ref{eq:welchextensionresult}) to hold is
that $S_1^2 = S_2^2 =...=S_m^2$.
Recently, Datta, Howard and Cochran have also proposed extensions to the Welch bound~\cite{DHC2009}.
\section{Packing and Covering Bounds for the RIP Ratio}
\label{sec:packing}
The motivation to derive another bound for the RIP ratio comes from the perspective that we can view the
CS matrix $\Phi$ as a collection of
column vectors in ${\mathbb R}^m$. We need to spread these vectors as far away from each other as possible
in order to ensure that the singular values of its $k$-column submatrices have good condition numbers.
Increasing the number of columns (i.e., increasing $n$) leads to crowding of these vectors in ${\mathbb R}^m$
that leads to a deterioration of the RIP ratio. We make these notions precise for $k=2$. For $k>2$, the
exact nature of the packing bounds are yet elusive.
As we saw from Theorem~\ref{theo:everymequalcols}, we need restrict our attention only to matrices of equi-normed columns.
Therefore,
the problem of designing good CS matrices for $k=2$ is equivalent to
finding arrangements of $n$ lines in ${\mathbb R}^m$ such that the minimum angle between the pairs of lines is maximized.
This problem has been studied extensively by Conway, Hardin, and Sloane \cite{SloaneGrassmannian}. Furthermore, converse and achievable bounds
can been derived by studying a related problem, namely of arrangements of $2n$ points on a Euclidean sphere
in ${\mathbb R}^m$. The latter problem has been studied independently by Chabauty, Shannon, and Wyner~\cite{Chabauty53,Shannon59,Wyner65}.
We state the main results that are relevant to our problem of bounding the RIP ratio. For a detailed description
and derivation of the relevant results from coding theory, see ~\cite{Ericson2001}.
\begin{DEFI}
The area of a {\em spherical cap} of radius $\beta$ on an $m$ dimensional Euclidean sphere of unit length is given by
\begin{equation*}
C_m(\beta) = k_{m} \int_0^\beta \sin^{m-2} \alpha d\alpha,
\end{equation*}
where $k_m$ is given by
\begin{equation*}
k_m=\frac {2 \pi ^ {(m-1)/2}} {\Gamma \left( \frac{m-1}{2} \right)}.
\end{equation*}
\end{DEFI}
Note that $2C_m(\pi)$ gives the surface area of a unit sphere in ${\mathbb R}^m$.
Packing the surface of the $m$-dimensional sphere using spherical caps gives rise to an upper bound on the
minimum angle $\theta$ that can be attained between the pairs of $2n$ points on the sphere. The upper bound
$\theta_{\max}$ is given by
$C_m \left(\frac{\theta_{\max}}{2} \right) = \frac{C_m(\pi)}{n}$.
Consequently, we obtain a lower bound on the RIP ratio, which is captured in Theorem~\ref{theo:packing} as
the packing bound. \\
{\bf Proof of Theorem~\ref{theo:packing}:}
Using (\ref{eq:optimalRd1d2}) to relate the minimum angular separation $\theta$ and the RIP ratio
$R(\Phi,2)$, we obtain
\begin{equation}
R(\Phi,2) = \frac{1+\cos(\theta)}{1-\cos(\theta)} = \cot^2 \left( \frac{\theta}{2} \right).
\end{equation}
The statement follows from the fact that for any arrangement of $2n$ points on the surface of the unit sphere in
${\mathbb R}^m$, the minimum angle between pairs of points is less than $\theta$ defined above, as a result of packing.
Note that the form of Theorem~\ref{theo:packing} is obtained by making a change of variable $\theta/2 \rightarrow \theta$.
\qed
Furthermore, {\em covering arguments} can be used to make a statement on achievability. Based on the results of Chabauty, Shannon, and
Wyner~\cite{Chabauty53,Shannon59,Wyner65,Ericson2001,Conway1998} we can guarantee the existence of
an arrangement of $2n$ points on the Euclidean sphere in ${\mathbb R}^m$
where the angular distance between every pair of points is
at least as large as $\theta$, where $C_m(\theta) = \frac{C_m(\pi)}{n}$.
Theorem~\ref{theo:cswachievable} captures the achievable bound on the RIP ratio obtained using the above arguments.
While we have successfully derived the structural bound for any $n$, $m$ and $k$,
the derivation of packing and covering bounds for $k>2$ remains an open problem. The main challenge is that
Theorem~\ref{theo:everymequalcols} cannot be extended beyond $k=2$. In fact, it is easy
to construct matrices where the RIP ratio for $k>2$ {\em increases} when we equalize the
column norms.
\begin{figure*}
\centering
\subfigure[$n=100$, $m=14$, $k=12$] {\epsfysize = 60mm \epsffile{histo_100_14_12.eps} \label{fig-histo_100_14_12_real}}
\subfigure[$n=100$, $m=20$, $k=12$] {\epsfysize = 60mm \epsffile{histo_100_20_12.eps} \label{fig-histo_100_20_12_real}} \\
\subfigure[$n=100$, $m=30$, $k=12$] {\epsfysize = 60mm \epsffile{histo_100_30_12.eps} \label{fig-histo_100_30_12_real}}
\subfigure[$n=100$, $m=60$, $k=12$] {\epsfysize = 60mm \epsffile{histo_100_60_12.eps} \label{fig-histo_100_60_12_real}}
\caption{\sl Histograms of the squared singular values of submatrices of $\Phi$. In each of the plots, we fix $n$, $m$ and $k$. We then plot,
in the same figure, the histograms for the $1$st, $4$th, $8$th and $12$th squared singular values $s_{p,1}^2$,
$s_{p,4}^2$, $s_{p,8}^2$, and $s_{p,12}^2$
of $25,000$ randomly chosen $m \times k$ sized submatrices of $\Phi$. The complete matrix $\Phi$
is chosen with $m$ prescribed singular values all equal to unity. The four red vertical lines in each plot
correspond to the values of the $1$st, $4$th, $8$th and $12$th zeros of the polynomial in (\ref{eq:thompson_result}).
Note the excellent match between the histogram peaks and the polynomial zeros; this observation suggests that the
zeros of the polynomial in (\ref{eq:thompson_result}) are in fact estimators (a mean, in some sense) of the squared singular values of
randomly chosen $m \times k$ submatrices of $\Phi$.}
\label{fig-histo}
\end{figure*}
\section{Relevance of the Structural Bounds in Statistical RIP}
\label{sec:stochrip}
It can be argued that the definition of RIP is too restrictive in the sense that the RIP constant
$\delta_k$ and the RIP ratio $R(\Phi,k)$ depend on extremal values of the singular values of
the submatrices. Because the number ${n \choose k}$ of submatrices of $\Phi$ is astronomical, it is very unlikely
that a randomly selected $k$-sparse signal has the exact sparsity pattern corresponding to the
submatrix of $\Phi$ with extreme values for the singular values. Rather than requiring every
$m \times k$-sized submatrix of $\Phi$ have their singular values bounded, we can conceive of a
{\em statistical RIP} where we allow a small fraction of the submatrices to have singular values outside of the bounds.
Along these lines, Tropp~\cite{Tropp2008_1,Tropp2008_2}, Calderbank, Howard, and Jafarpour~\cite{CalderbankDetRIP} and Gurevich and Hadani~\cite{gurevich}
have proposed the notion of Statistical RIP.
We demonstrate how the parameters of the structural bound play an important role in capturing the estimates of
singular values in a randomly chosen $m \times k$-sized submatrix of $\Phi$. The motivation comes from the fact that
the quantities $r_1^2$ and $r_k^2$ (of Theorems ~\ref{theo:main1} and \ref{theo:thompsonrip})
in some sense capture the mean of the squared singular values $s_{p,1}^2$ and $s_{p,k}^2$, respectively.
In fact, while proving Theorems ~\ref{theo:main1} and \ref{theo:thompsonrip}, we have shown that if
the $s_{p,1}^2$'s are not all equal, then
some of the $s_{p,1}^2$'s lie to the left of the real number line from $r_1^2$ and some lie to the right of $r_1^2$.
Similarly for the set of $s_{p,k}^2$ and $r_k^2$. This leads one to wonder if
in fact $r_i^2$ is in some sense, the mean of $s_{p,i}^2$. In other words, is it possible that the zeros of the polynomial in (\ref{eq:thompson_result}) are in fact estimates of the
squared singular values of a randomly selected submatrix of $\Phi$?
We ran the following simulation to test our intuition. We picked a
single $\Phi$ of moderate size $m \times n$ with a prescribed set of singular values (all ones)
and randomly selected a large number of submatrices of $\Phi$
of size $m \times k$. We computed the $k$ singular values of each of the selected submatrices.
We plot the $k$ histograms
of the respective squared singular values and compare them against the $r_i^2$'s. Figure ~\ref{fig-histo}
shows these plots for a set of $4$ out of $k$ singular values (we chose only $4$ in order to prevent clutter in the plot).
Note that the $r_i^2$'s provide remarkably good estimates for the $s_{p,i}^2$'s. This observation strongly
suggests that the
roots of the polynomial in (\ref{eq:thompson_result}) play a crucial role in statistical RIP, irrespective of which of the two converse bounds (structural
or packing) is the tighter deterministic bound. We believe that this observation serves as a starting point for
analysis in statistical RIP.
\section{Conclusions}
\label{sec:conclusions}
In this paper,
we have derived two deterministic converse
bounds for RIP ratio. The first bound is based on structural bounds for singular values of submatrices and the second
bound is based on packing arguments. We have also derived a deterministic achievable bound on RIP ratio
using covering arguments. The derivation of the three bounds offer
rich geometric interpretation and illuminate the relationships between CS
matrices and equi-angular tight frames,
codes on Grassmannian spaces and Euclidean spheres, and the Generalized
Pythagorean Theorem.
A summary of our key results is given below:
\begin{enumerate}
\item There is a large gap between the RIP ratio of Gaussian matrices and the achievable
bounds. This observation points to the existence of CS matrices that are far superior than Gaussian
matrices in terms of the RIP ratio.
\item For small values of $n$, the structural bound is the tighter of the two converse
bounds, whereas for large values of $n$, the packing bound is tighter.
\item We compared the three bounds to the
RIP ratio of Gaussian matrices, as well as the best known matrix for the given problem size. We used
the results of Conway, Hardin, and Sloane \cite{SloaneGrassmannian,SloaneGrassmannianWeb,SloaneGrassmannianWebTable}, who have run extensive computer simulations
to extract the best known packings in Grassmannian spaces.
\item While the structural bound for the RIP ratio has been derived for any $n$, $m$, and $k$, we presently
have the packing and covering bounds only for $k=2$. We believe that the result for $k=2$ establishes a
starting point to investigate the packing and covering bounds for $k>2$.
\item The parameters of the structural bound can shed light on the statistical
RIP that was proposed recently~\cite{CalderbankDetRIP,gurevich}. In particular, we demonstrated a way to estimate the singular
values of a randomly chosen submatrix of $\Phi$.
\item We showed that the structural bound for $k=2$ is equivalent to the Welch bound~\cite{WelchEquiangular}.
We have used the structural bound for $k>3$ to extend the Welch bound to higher orders.
\end{enumerate}
The present study of deterministic bounds for RIP opens up many interesting research questions.
While we have derived deterministic RIP ratio bounds that apply to all matrices in ${\mathbb R}^{m \times n}$,
it would be valuable to derive the deterministic RIP ratio bounds
for special class of CS matrices within ${\mathbb R}^{m \times n}$,
such as $\{-1,1\}$ matrices, sparse matrices,
and matrices that have block zeroes that appear in Distributed Compressed Sensing~\cite{DCS}.
Analysis of these special class of matrices would also help measure the penalty in terms of the
increase in RIP ratio we need to tolerate.
Next, we plan to characterize the exact relationship between the stochastic RIP described in Section~\ref{sec:stochrip} and
the parameters of the structural bound.
Finally, packing and covering bounds for $k>2$ remains an open problem.
|
\section{Introduction}
\label{sec1}
A fundamental postulate in Einstein's theory of gravitation is that physics only depends on intrinsic geometric properties of curved space-time. Many approaches to gravity and cosmology assume that space-time is a brane, i.e., a submanifold embedded in a higher dimensional ambient space \cite{Maartens} (see \cite{Branes} for an extensive list of commented references). In such a brane world scenario, Einstein's postulate cannot be assumed but rather is a property that should be derived from an underlying theory. One intriguing possibility is that this property holds true only in some approximation. In this case one might be able to predict physical effects that depend on extrinsic curvature. This would be interesting as a means to falsify theoretical proposals. It also might suggest experiments that give direct evidence for (or against) the existence of extra dimensions. In this paper we propose a possible cause for extrinsic curvature effects that, as we believe, is potentially relevant in brane world scenarios. We also discuss possible physical implications. Our proposal is motivated by well-established results on constrained quantum mechanics, as discussed in the next paragraph.
As proposed already by Schr\"odinger in 1926, it is natural to postulate that the quantum mechanical Hamiltonian for a free particle on a submanifold of three dimensional Euclidean space is proportional to the Laplace-Beltrami operator on this submanifold, and it thus only depends on intrinsic geometry; see Equation~(31) in \cite{Schrodinger}. A more physical method is to derive this Hamiltonian as follows: use the well-established Hamiltonian in three dimensional space and restrict the particle to the manifold by a strong confining potential. As discovered by Jensen and Koppe \cite{JensenKoppe} in this context,\footnote{A similar result was found earlier by Marcus \cite{Marcus} in an effective quantum model for chemical reactions.} the effective Hamiltonian thus obtained contains, in addition to the expected Laplace-Beltrami operator term, an induced potential that depends on intrinsic- and extrinsic curvature; see \cite{C1,C2,MD,Mitchell,FH,SJ,Gol,WT} for generalizations and alternative derivations of this result. For example, the induced potential for a curve is attractive and proportional to the curvature. This explains why, as shown by Exner and Seba \cite{ES}, an electron on a wire has bound states when the wire has non-zero curvature;\footnote{Note that the intrinsic geometry of a curve is trivial, and thus Schr\"odinger's postulated effective Hamiltonian cannot explain such bound states.} see also \cite{GJ} for similar results in the context of electromagnetic waveguides. The existence of the induced potential was confirmed experimentally; see e.g.\ \cite{CLMTY} and references therein.
Consider a real-valued Klein-Gordon field $\phi$ on curved space-time embedded in higher dimensional Minkowski space. We propose to {\em derive} an effective equation of motion for $\phi$ from a Klein-Gordon field on ambient space restricted to the space-time submanifold by a strong confining potential. As we show, this yields an effective description by the following modified Klein-Gordon equation on space-time\footnote{Our notation is explained in the next section. }
\begin{equation}
\label{KG1}
\Bigl( \hbar^2\Bigl[ |g|^{-1/2} \partial_\mu|g|^{1/2}g^{\mu\nu}\partial_\nu + V_{ind}(x)\Bigr]+ (m_0 c)^2 \Bigr)\phi(x) =0
\end{equation}
with the induced potential $V_{ind}$ in Equation~\Ref{Vind} below (we use Planck's constant $\hbar$ and the vacuum velocity of light $c$ only here, to emphasize that the induced potential is a correction to the Laplace-Beltrami operator; in the rest of the paper we set $\hbar=c=1$). In fact, our result is more general: we use the Lagrangian formalism, and \Ref{KG1} corresponds to the Euler-Lagrange equations of the non-interacting part of the action only, but our result applies also to the case with interactions. It is worth noting that $V_{ind}$ is universal in the sense that it is independent of the details of the confining potential. We also investigate if and when $V_{ind}$ could lead to measurable effects. For that we compute $V_{ind}$ for various special cases.
We emphasize that, for given space-time, the embedding in a higher dimensional Minkowski space is not unique, and the induced potential depends on the embedding. Our proposal can therefore only lead to a definite prediction in combination with a theory that fixes the embedding. We do not assume such a theory but instead use simple embeddings in flat ambient spaces with the lowest possible dimension. Fortunately, in many examples of physical interest (including the ones we discuss), there exists one such embedding which is preferred by its naturalness and simplicity; see \cite{Rosen} for a list of known examples. We thus believe that our approach is justified by "Ockham's razor". The investigation of more complicated embeddings (motivated by string theory etc.) is left to future work.
While the formula for the induced potential we obtain is identical with the obvious generalization of the well-known one in constrained quantum mechanics \cite{JensenKoppe,C2,MD} from Riemannian- to pseudo-Riemannian spaces, the physics is very different, and the derivations of the induced potential in the literature do not apply to our case. We therefore present a different derivation for pseudo-Riemannian manifolds $\mathcal{M}^{q,n}$ of arbitrary signatures $(q,n)$ embedded in flat ambient spaces $\mathbb{R}^{q,n+p}$ (i.e., the number $q\geq 0$ and $n\geq 1$ of time- and space-like dimensions in $\mathcal{M}^{q,n}$ is arbitrary, and so is the number $p\geq 1$ of the extra dimensions). To not further burden our notation, we restrict our derivation to the case $q=1$ of main interest to us, but the generalization to arbitrary $q$ is obvious. To be more specific: Derivations of the induced potential in constrained quantum mechanics are usually based on equations of motion; see \cite{WT}, Section~I.A for a more detailed discussion of the history and the different levels of mathematical rigor of these derivations. A key point in such quantum mechanical derivations is the probability interpretation of the quantum mechanical wave function. This provides an argument to use a scaling factor for the wave function in ambient space, and this scaling factor leads to the induced potential; see e.g.\ \cite{SJ}, Equation~(12) {\em ff} for a lucid discussion of this point. However, Klein-Gordon fields do not have such a probability interpretation, and this argument therefore cannot be used. Instead, we start with the standard action for a Klein-Gordon field on flat ambient space and with a suitable strong confining potential; see Equation~\Ref{S0} below. By expanding this field in suitable modes and straightforward computations, we find that this action describes coupled Klein-Gordon fields on space-time; see Equation~\Ref{action2}. Using a standard physics argument, we finally reduce the latter to an effective action for a single Klein-Gordon fields; see Equation~\Ref{Seff}. The above-mentioned scaling factor arises in this derivation for purely mathematical reasons. Our derivation suggests that such an effective action cannot be obtained in cases where the number of time-like directions in physical space-time and ambient space are different.
In the rest of this paper we first introduce our notation and present our derivation of the induced potential for confined Klein-Gordon fields (Section~\ref{sec2}). Our examples are in Sections~\ref{sec3}--\ref{sec6}: We first give the induced potential for the Schwarzschild metric and discuss its possible physical implications for black holes (Section~\ref{sec3}). We then consider the induced potential for the closed Robertson-Walker metric (Section~\ref{sec5}), and we use this to construct and study a brane-world model of cosmological inflation taking into account extrinsic curvature effects (Section~\ref{sec6}). We end with conclusions in Section~\ref{sec7}. Some details of our computations are given in two appendices.
We will mention basic results on general relativity and cosmology that are discussed in several textbook (including \cite{Borner,LythLiddle09}) without further reference.
\section{General result}
\label{sec2}
We assume that space-time $\mathcal{M}\equiv \mathcal{M}^{1,n}$ is a $(n+1)$-dimensional Lorentzian manifold embedded in $(n+p+1)$-dimensional Minkowski space ${\mathbb R}^{1,n+p}$.
\noindent {\bf Notation:} We use capital latin letters $M,N,\dots$ for indices running over $0,1,2,\dots,n+p$, and denote as $\mathbf{Z}\equiv(Z^0,Z^1,Z^2,\dots,Z^{n+p})$ inertial coordinates in ${\mathbb R}^{1,n+p}$, i.e., the line element in these coordinates is (the following defines our sign conventions)
\begin{equation}
\label{ds2}
ds^2 = \eta_{MN}dZ^MdZ^N \equiv (dZ^0)^2 - (dZ^1)^2 - \dots - (dZ^{n+p})^2.
\end{equation}
Greek letters $\mu,\nu,\lambda,\sigma,\ldots$ and lower case latin letters $i,j,k\dots$ are used for indices running over $0,1,2,\dots,n$ and $n+1,n+2,\dots,n+p$, respectively. We write bold face letters for vectors in ${\mathbb R}^{1,n+p}$ and a dot for the scalar product of such vectors, i.e., $\mathbf{v}\cdot\mathbf{w}\equiv\eta_{MN} v^M w^N$. We assume that space-time $\mathcal{M}$ can be (locally) parametrized by a function $\mathbf{Z} = \mathbf{f}(x)$ with $x\equiv(x^0,\ldots,x^n)$ in an open subset of $\mathbb{R}^{1,n}$. The metric tensor of $\mathcal{M}$ is then $g_{\mu\nu}=\mathbf{t}_\mu \cdot \mathbf{t}_\nu$, with tangent vectors $\mathbf{t}_\mu(x) \equiv \partial_\mu\mathbf{f}(x)$.\footnote{We write $\partial_M$, $\partial_\mu$ and $\partial_i$ short for $\frac{\partial}{\partial Z^M}$, $\frac{\partial}{\partial x^\mu}$ and $\frac{\partial}{\partial y^i}$, respectively. We always assume implicitly that functions we introduce are differentiable up to the degrees needed.} Using $y\equiv (y^{n+1},\dots,y^{n+p}) \in \mathbb{R}^{p}$ we construct a local coordinate system of an open neighborhood of $\mathcal{M} \subset \mathbb{R}^{1,n+p}$ by setting
\begin{equation}
\label{tildeZ}
\mathbf{\tilde{f}}(x,y) = \mathbf{f}(x) + y^i \mathbf{n}_i(x)
\end{equation}
with $p$ linearly independent space-like vectors $\mathbf{n}_i$ orthogonal to all tangent vectors
and with constant scalar products, i.e.,
\begin{equation}
\label{t and n}
\mathbf{t}_\mu (x) \cdot \mathbf{n}_i (x) =0,\quad \mathbf{n}_i (x) \cdot \mathbf{n}_j (x) = -h_{ij}
\end{equation}
for all $x$, $\mu$ and $i$, with constant $h_{ij}$ defined by this equation. We can further restrict the vectors $\mathbf{n}_i$ by the additional condition $h_{ij}=\delta_{ij}$, without loss of generality. However, we sometimes find it convenient to leave $h_{ij}$ general in our equations. We assume that there exists $\epsilon>0$ such that the coordinate system $(x,y)$ is one-to-one for all $y$ such that $|y|\equiv \sqrt{h_{ij}y^iy^j}<\epsilon$ (as discussed in the beginning of Appendix~\ref{appA1}, we believe that this is a minor restriction). We denote the metric tensor on ${\mathbb R}^{1,n+p}$ in the coordinates $(x,y)$ by $G_{MN}(x,y)$. Here and in the following we write $|h|$, $|g|$ and $|G|$ for the absolute values of the determinants of $(h_{ij})$, $(g_{\mu\nu})$ and $(G_{MN})$, respectively. We define $G^{MN}$ as usual: $G^{ML}G_{LN}=\delta^M_{\phantom M N}$, and similarly for $g^{\mu\nu}$ and $h^{ij}$.
We consider a Klein-Gordon field $\Phi\equiv \Phi(\mathbf{Z})$ on ambient space ${\mathbb R}^{1,n+p}$ with the usual dynamics but confined to the region close to the submanifold $\mathcal{M}$ by a potential $V_{conf}(\mathbf{Z})$, i.e., the (free part of) the action is
\begin{equation}
\label{S0}
S_0=\frac12\int_{{\mathbb R}^{1,n+p}} d^{n+p+1}Z\Bigl(\eta^{MN}\bigl(\partial_M \Phi\bigr)\bigl(\partial_N\Phi\bigr) -(m_{bare}^2+V_{conf})\Phi^2 \Bigr)
\end{equation}
with the ``bare mass parameter'' $m_{bare}^2\in{\mathbb R}$. We assume that the potential $V_{conf}$ is strongly confining to $\mathcal{M}$, i.e., in the coordinates $(x,y)$ it has the form
\begin{equation}
\label{Vconf}
\tilde V_{conf}(x,y)\define
V_{conf}(\tilde{\mathbf{f}}(x,y))=\frac1{\epsilon^2} V(y/\epsilon)+ O(|y|)
\end{equation}
with $\epsilon>0$ a convenient scaling parameter assumed to be small;\footnote{Introducing this scaling parameter is a useful mathematical trick allowing to cleanly separate different energy scales; see e.g.\ \cite{WT} for a lucid discussion of this point.} we allow for a possible correction term $O(|y|)$ depending on $x$ but vanishing at least linearly with $|y|$ as $|y|\to 0$. We assume that $V(y)$ is such that the eigenvalue equation\footnote{To be more precise, we assume that $V(y)$ is such that the following is an eigenvalue equation of a self-adjoint Schr\"odinger operator on the Hilbert space of square integrable function on ${\mathbb R}^p$.}
\begin{equation}
\label{def chi}
-h^{ij}\partial_i\partial_j\chi_\alpha(y) + V(y)\chi_\alpha(y)=\mu_\alpha \chi_\alpha(y)
\end{equation}
has a unique solution $\chi_0$ corresponding to the smallest possible eigenvalue $\mu_0$ and such that $\int d^py\, |h|^{1/2}|y|^n|\chi_0(y)|^2$ is finite for all $n=0,1,2,\ldots$, and $\mu_\alpha-\mu_0>0$ for all $\alpha\neq 0$; here and in the following, indices $\alpha,\beta\ldots$ denote quantum numbers labeling the solutions of the eigenvalue equation in \Ref{def chi}. Below we refer to such a potential $V(y)$ as {\em suitable}. We choose the eigenfunctions $\chi_\alpha$ to be real-valued and normalized such that
\begin{equation}
\label{def chi 1}
\int d^p y\, |h|^{1/2}\chi_\alpha(y)\chi_\beta(y)=\delta_{\alpha\beta}.
\end{equation}
To be specific we mention one example for a suitable potential allowing for simple computations of all solutions of \Ref{def chi}: $h_{ij}=\delta_{ij}$ and
\begin{equation}
\label{box}
V(y) = \lim_{v_0\to+\infty}v_0\Bigl(1 - \prod_{j=n+1}^{n+p}\theta(\ell-|y^j|) \Bigr)
\end{equation}
with $\ell>0$ and the Heaviside function $\theta$ (``infinite box potential'') but, as discussed in Appendix~\ref{appA1}, the same result is obtained for a large class of potentials. In particular, $V$ can be bounded and such that the eigenvalue equation in \Ref{def chi} also has scattering solutions (in this case the symbols $\sum_\alpha$ and $\delta_{\alpha\beta}$ have to be partly interpreted as integral and Dirac delta).
One can expand $\tilde\Phi(x,y)\define\Phi(\mathbf{\tilde f}(x,y))$ in the eigenfunctions $\epsilon^{-p/2}\chi_\alpha(y/\epsilon)$ and thus rewrite the action in \Ref{S0} as an action of an infinite number of fields $\phi_\alpha(x)$ on space-time $\mathcal{M}$. As explained in Appendix~\ref{appA2}, a key point in this computation is that, to correct a mismatch of Jacobian determinants, one has to include a scaling factor $(|h||g|/|G|)^{1/4}$ in this expansion as follows,
\begin{equation}
\label{expansion}
\tilde\Phi(x,y)=\left(\frac{|h||g(x)|}{|G(x,y)|}\right)^{1/4}\sum_\alpha\phi_\alpha(x) \epsilon^{-p/2}\chi_\alpha(y/\epsilon).
\end{equation}
By a straightforward computation we obtain the following.
\noindent \textbf{Result:} {\em For suitable confining potentials $V$ and sufficiently small $\epsilon>0$, the action in \Ref{S0} equals
\begin{equation}
\label{action2}
S_0 = \frac12\int_{\mathcal{M}} d^{n+1}x\, |g|^{1/2}\sum_\alpha \Biggl(g^{\mu\nu} \bigl(D_\mu\phi\bigr)_\alpha\bigl(D_\nu \phi\bigr)_\alpha - \bigl(m_{bare}^2 + \mu_\alpha/\epsilon^2+ V_{ind}\bigr)\phi_\alpha^2 \Biggr) + O(\epsilon)
\end{equation}
\begin{equation}
\label{Ddef}
\bigl(D_\mu\phi\bigr)_\alpha\equiv \partial_\mu\phi_\alpha + \sum_{\beta\neq \alpha }C_{\mu\alpha\beta}\phi_\beta
\end{equation}
with
\begin{equation}
\label{Cdef}
C_{\mu\alpha\beta} = A_{j\mu}^{\phantom{j\nu}i} \int_{{\mathbb R}^p} d^py\, |h|^{1/2} y^j \chi_\beta (y) \partial_i \chi_\alpha (y)
\end{equation}
and
\begin{equation}
\begin{split}
\label{Vind}
V_{ind} = \frac{1}{4} h^{ij}\left(\alpha_{i\lambda}^{\phantom{i\lambda}\lambda}\alpha_{j\sigma}^{\phantom{j\sigma}\sigma}- 2\alpha_{i\lambda}^{\phantom{i\lambda}\sigma}\alpha_{j\sigma}^{\phantom{j\sigma}\lambda}\right)
\end{split}
\end{equation}
where $\alpha_{i\mu}^{\phantom{i\mu}\lambda}(x)$ and $A_{i\mu}^{\phantom{i\mu}k}(x)$ are defined by the following equation,
\begin{equation}
\label{alpha and A}
\partial_\mu \mathbf{n}_i = - \alpha_{i \mu}^{\phantom{i\mu}\lambda} \mathbf{t}_\lambda - A_{i\mu}^{\phantom{i\mu}k} \mathbf{n}_k.
\end{equation}
}
\noindent (See Appendix~\ref{appA} for a derivation of this result.)
We note in passing that the action in \Ref{action2}--\Ref{Vind} has a gauge theory structure discussed in \cite{MD,SJ}, for example, but this is not used in the present paper.
Thus, up to terms $O(\epsilon)$ vanishing in the limit $\epsilon\to 0^+$, the confined Klein-Gordon action on ambient space in \Ref{S0} is equivalent to an action of Klein-Gordon fields $\phi_\alpha$ on $\mathcal{M}$ with effective masses
\begin{equation}
m_\alpha^2 = m_{bare}^2 + \mu_\alpha/\epsilon^2
\end{equation}
and additional potential- and derivative terms. It is natural to fix the renormalized mass $m_0^2$ and choose $m_{bare}^2=m_0^2-\mu_0/\epsilon^2$. The other mass parameters $m^2_{\alpha\neq 0}=m_0^2+(\mu_\alpha-\mu_0)/\epsilon^2$ are then much larger than $m_0^2$. Standard physics arguments suggest that, for sufficiently small values of $\epsilon$, only the Klein-Gordon field $\phi_0$ with the ``small'' mass parameter $m_0^2$ is relevant for the low energy physics properties of the model. It therefore is a good approximation to simplify the model by replacing the action in \Ref{action2} with
\begin{equation}
\label{Seff}
S^{(0)}_{eff} = \frac12\int_{\mathcal{M}} d^{n+1}x\, |g|^{1/2} \left(g^{\mu\nu} \bigl( \partial_\mu \phi\bigr)\bigl(\partial_\nu \phi\bigr) - (m^2_0 + V_{ind}) \phi^2 \right)
\end{equation}
and $\phi\equiv \phi_0$. This is our result described in the introduction. It is important that it is independent of the details of the confining mechanism: changing $V$ can be compensated by a change of the bare mass parameter $m_{bare}^2$ and thus is irrelevant.
As already mentioned, the generalization of this result to interacting Klein-Gordon fields is straightforward. In particular, changing the action in \Ref{S0} by a $\Phi^4$-interaction term,
\begin{equation}
S = S_0 - \frac{\lambda_{bare}}4\int_{{\mathbb R}^{r+p+1}}d^{n+p+1}Z\, \Phi^4,
\end{equation}
leads to the following change of the effective action in \Ref{Seff},
\begin{equation}
\label{Seff1}
S_{eff} = S^{(0)}_{eff} - \frac{\lambda_0}4 \int_{\mathcal{M}} d^{n+1}x\, |g|^{1/2}\phi^4
\end{equation}
with the renormalized interaction strength
\begin{equation}
\lambda_0 = \lambda_{bare}\epsilon^{-p} \int_{{\mathbb R}^p} d^p y\, |h|^{1/2}\chi_0(y)^4.
\end{equation}
Again it is natural to fix $\lambda_0>0$ and adapt $\lambda_{bare}$ accordingly, i.e., the effective action is independent of the details of the confining mechanism also in the presence of interactions.
\section{The Schwarzschild black hole}
\label{sec3}
The Schwarzschild metric is a spherically symmetric solution of Einstein's equations describing a black hole with mass $M$ in an otherwise empty universe. It is given by the line element
\begin{equation}
\label{ds2_Schwarzschild}
ds^2 = \left( 1 - \frac{r_s}{r}\right) dt^2 - \left( 1 - \frac{r_s}{r}\right)^{-1} dr^2 - r^2 (d\theta^2 + \sin^2(\theta)d\varphi^2)
\end{equation}
where $r_s = 2GM$ is the Schwarzschild radius and $x=(t,r,\theta,\varphi)$ the Schwarzschild coordinates, as usual.
As proved by Kasner \cite{K1}, the Minkowski space-time of lowest dimension allowing an embedding of Schwarzschild space-time is $\mathbb{R}^{1,5}$. One such embedding found by Fronsdal \cite{Fronsdal} is given by
\begin{equation}
\begin{split}
\label{embed_Schwarzschild}
&Z^0 = \begin{cases} 2r_s(1-r_s/r)^{1/2} \sinh(t/[2r_s])& (r>r_s)\\ 2r_s(r_s/r-1)^{1/2} \cosh(t/[2r_s])& (0<r<r_s) \end{cases}\\
&Z^1 = \begin{cases} 2r_s(1-r_s/r)^{1/2} \cosh(t/[2r_s])& (r>r_s)\\ 2r_s(r_s/r-1)^{1/2} \sinh(t/[2r_s])& (0<r<r_s) \end{cases}\\
&Z^2 = g(r)\\
&Z^3 = r\sin(\theta)\cos(\varphi)\\
&Z^4 = r\sin(\theta)\sin(\varphi)\\
&Z^5 = r\cos(\theta)
\end{split}
\end{equation}
where
\begin{equation}
\label{gprime}
g'(r)\equiv \frac{dg(r)}{dr} = \sqrt{\frac{r_s(r^2+rr_s+r_s^2)}{r^3}}.
\end{equation}
This embedding is natural in that no other complete embedding of the Schwarzschild solution in the six-dimensional Minkowski space-time exists \cite{Sassi}.
By straightforward computations we find the induced potential (see Appendix~\ref{appB1} for details)
\begin{equation}
\label{Schwarzschild_result1}
V_{ind}(r) = -\frac{(\hat r^3+ \hat r^2+ \hat r+9)(\hat r^2+1)(\hat r+1)}{16r_s^2 \hat r^4(\hat r^2+\hat r+1)},\quad \hat r\equiv r/r_s
\end{equation}
which is non-singular in the whole region $0<r<\infty$. Considering that the details of the embedding are different for $r>r_s$ and $r<r_s$, it is remarkable that the induced potential is continuous at $r=r_s$. Note that $V_{ind}(r)$ is strictly monotone and has the following asymptotic behavior,
\begin{equation}
\label{Schwarzschild_result2}
V_{ind}(r) =\begin{cases} -9/(16 r_s^2 \hat r^4)\Bigl( 1 + \hat r/9 +O(\hat r^2)\Bigr) & (r\to 0^+)\\
-1/r_s^2\Bigl(1-3(\hat r-1)+O\bigl((\hat r-1)^2\bigr)\Bigr)& (r\to r_s) \\
-1/(4r_s)^2\Bigl( 1 + 1/\hat r + O(1/\hat r^{2}) \Bigr) & (r\to \infty) \end{cases}.
\end{equation}
Moreover, it approaches its limiting value $-1/(4r_s)^2$ rather rapidly and, for distances $r$ larger than $100r_s$, $V_{ind}(r)$ differs from its limiting value by less than $1\%$. Note that the magnitude of this limiting value corresponds to a boson mass, in SI units,
\begin{equation}
\label{hbar}
\frac{\hbar}{4 c r_s} \approx \frac{3.31\times 10^{10}\, \mathrm{kg}}{M}\frac{\mathrm{GeV}}{c^2}.
\end{equation}
We conclude this section with a short discussion of the possible physical relevance of our result. It is worth noting that the boson-mass equivalent of the magnitude of the induced potential in \Ref{hbar} is proportional to the {\em inverse} of the black hole mass $M$. Moreover, it seems that only very light black holes, with masses $10^{10}\,$kg or smaller, could give rise to induced potentials that have measurable effects, and the spatial variation of this potential is significant only very close to such a black hole. Thus the natural candidates to search for such effects are {\em primordial black holes}; see e.g.\ \cite{Carr} and references therein. It is remarkable that, very far from a black hole, the induced potential renormalizes boson masses by a negative constant. If many primordial black holes exist in the universe their cumulative mass renormalization effect could be quite large.
\section{Robertson-Walker space-time}
\label{sec5}
The Robertson-Walker space-time describes the evolution of a homogenous isotropic universe. Its metric depends on a real parameter $K$ whose sign determines if the universe is open ($K>0$), flat ($K=0$), or closed ($K<0$). All cases can be described by a line element of the form
\begin{equation}
\label{ds2_RW}
ds^2 = dt^2 - a(t)^2 \left( dr^2 + S(r)^2(d\theta^2 + \sin^2(\theta) d\varphi^2) \right)
\end{equation}
in coordinates $x=(t,r,\theta,\varphi)$, with $t$ the cosmological time, $a(t)$ the so-called {\em scale factor}, and
\begin{equation}
S(r)\define \begin{cases} \sin(\sqrt{K}r)/\sqrt{K}& (K>0)\\ r& (K=0)\\ \sinh(\sqrt{|K|}r)/\sqrt{|K|} & (K<0)\end{cases} .
\end{equation}
This space-time can be naturally\footnote{We believe that there is no other embedding in five dimensional Minkowski space-time, but we are not aware of a proof of this in the literature.} embedded in $\mathbb{R}^{1,4}$ as follows (see \cite{Rosen}, B.5.),
\begin{equation}
\label{embed_RW}
\begin{split}
&Z^0 = \begin{cases} b(t)/\sqrt{K}& (K>0)\\ (1/2)\bigl( r^2/r_0+ r_0\bigr)a(t) + B(t) & (K=0)\\ a(t)C(r) & (K<0) \end{cases} \\
&Z^1 = \begin{cases} a(t) C(r) & (K>0)\\ (1/2)\bigl( r^2/r_0 - r_0\bigr)a(t) + B(t) & (K=0)\\ b(t)/\sqrt{|K|} & (K<0) \end{cases} \\
&Z^2 = a(t) S(r) \sin(\theta) \cos(\varphi)\\
&Z^3 = a(t) S(r) \sin(\theta) \sin(\varphi)\\
&Z^4 = a(t) S(r) \cos(\theta)
\end{split}
\end{equation}
with functions $b$ and $B$ defined by the following equations,
\begin{equation}
\label{bdot}
\dot{b}(t) = \sqrt{K+\dot a(t)^2} \quad (K\neq 0)
\end{equation}
and
\begin{equation}
\dot B(t) = \frac1{2r_0 \dot a(t)}
\end{equation}
($\dot b(t)\equiv db(t)/dt$ etc.), and
\begin{equation}
C(r)\define \begin{cases} \cos(\sqrt{K} r)/\sqrt{K} & (K>0)\\ \cosh(\sqrt{|K|} r)/\sqrt{|K|} & (K<0)\end{cases} ;
\end{equation}
$r_0>0$ is an arbitrary parameter.
By straightforward computations we find the induced potential (see Appendix~\ref{appB3} for details)
\begin{equation}
\label{RWpot}
V_{ind} = \frac{1}{4} \left( 6\frac{\ddot{a}}{a} + 3\frac{K+\dot{a}^2}{a^2} - \frac{\ddot{a}^2}{K+\dot{a}^2} \right)
\end{equation}
for arbitrary $K$. Considering that the details of the embedding are different in the three cases $K>0$, $K=0$, and $K<0$, it is remarkable that the induced potential has such a simple form, and, in particular, that it is continuous at $K=0$.
We conclude this section with a preliminary discussion of the possible relevance of this induced potential in cosmology. To simplify some formulas we only consider the case $K=0$. Inserting $H\define \dot a/a$ we find by simple computations
\begin{equation}
V_{ind} = -\frac{H^2}{4}\Bigl( q^2+6q-3\Bigr)
\end{equation}
with $q\equiv \ddot a a/\dot a^2=-(\dot H+H^2)$ the usual {\em deceleration parameter}. Thus $V_{ind}\geq 0$ if the deceleration parameter is in the range $-6.46(4)\leq q\leq 0.464(1)$ and $<0$ otherwise, and $V_{ind}$ has typically the same order of magnitude as $H^2$. Since the value of the Hubble constant $H_0$ today corresponds to a boson mass of approximately $10^{-33}\,$eV, this suggests to us that the induced potential is negligibly small in the universe at present and far back in time. To see if the induced potential could have an effect at early times we compute $V_{ind}$ for a universe with a scale factor growing as $a(t)=(t/t_*)^x$ for some exponent $x>0$ and some constant $t_*$, with $t$ the time after the big bang at $t=0$. We find
\begin{equation}
V_{ind}=\frac{8x^2-4x-1}{4t^2},
\end{equation}
i.e., as $t\to 0^+$, the induced potential diverges towards $-\infty$ for $0\leq x<(1+\sqrt{3})/4$ and towards $+\infty$ for $x>(1+\sqrt{3})/4$. This suggest that the induced potential could have had an important effect in the early universe. In the next section we propose and study a self-contained model for the evolution of the early universe, taking into account the induced potential. As we will see, this model predicts that, as $t\to 0^+$, the scaling factor can vanish like $a(t)\to (t/t_*)^x$ with $x=(1+\sqrt{3})/4$ the critical value of the exponent, and such behavior is impossible without the induced potential.
\section{Extrinsic curvature effects in the early universe}
\label{sec6}
Models of the early universe often assume a Robertson-Walker metric and matter represented by an isotropic Klein-Gordon field; see e.g.\ \cite{Linde,LythRiotto} for reviews or \cite{LythLiddle09} for a recent textbook on this topic. In this section we propose a generalization of such a model taking into account extrinsic curvature effects. We also present results on the solution of this model and, to put them in perspective, compare them with results for the corresponding standard model where the external curvature effects are ignored.
We obtain this model by adding our induced potential term to the action consisting of the usual Einstein-Hilbert term, cosmological term, and Klein-Gordon term with $\phi^4$-interaction \cite{LythLiddle09}:\footnote{In this section we write $\lambda$ and $m$ short for $\lambda_0$ and $m_0$, respectively. Moreover, the meanings of the symbols $x$, $y$, and $\phi_n$ are different from the ones in Section~\ref{sec2}.}
\begin{equation}
\label{S}
S= \frac12\int d^4 x\, |g|^{1/2}\Bigl( M_{Pl}^2(R-2\Lambda) + g^{\mu\nu}(\partial_\mu\phi)(\partial_\nu\phi) - (m^2+ \chi V_{ind})\phi^2 -\frac{\lambda}2\phi^4\Bigr)
\end{equation}
with the Ricci scalar $R$ and the induced potential $V_{ind}$ in \Ref{Vind}. Our model parameters are the reduced Planck mass $M_{Pl}=1/\sqrt{8\pi G}$, the cosmological constant $\Lambda$, and the renormalized mass $m$ and coupling constant $\lambda$ of the Klein-Gordon field, respectively. To avoid writing similar formulas twice we use a parameter $\chi$ which is either $0$ (for the standard model case) or $1$ (for the extended model case with extrinsic curvature effects included).
\subsection{Model details}
\label{RWmodels}
We are interested in a homogeneous Klein-Gordon field $\phi=\phi(t)$ (depending only on cosmological time $t$) on Robertson-Walker space-time with scaling factor $a(t)$. The standard method to derive differential equations determining the time evolution of $a$ and $\phi$ in the case $\chi=0$ is as follows: restrict the Euler-Lagrange equations obtained from the action in \Ref{S} to the Robertson-Walker metric in \Ref{ds2_RW} and Klein-Gordon fields $\phi=\phi(t)$. Unfortunately we cannot use this method in the case $\chi=1$: this would require a formula describing how the induced potential changes with arbitrary variations of the metric, but this we do not have (since the induced potential depends on the embedding, and we do not know how to naturally change the embedding with the metric in general).
We thus use an alternative method which, in the standard case $\chi=0$, leads to the same result as the method just described: We insert the metric in \Ref{ds2_RW} and $\phi=\phi(t)$ into the action in \Ref{S}. This yields $S=const \int dt\, L$ (we can ignore the multiplicative constant) with the Lagrangian
\begin{equation}
\label{L}
\begin{split}
L= M_{Pl}^2(-3a\dot{a}^2-\Lambda a^3) + \frac{a^3}2\Biggl(\dot\phi^2- \Bigl(m^2 + \frac{\chi}{4} \Bigl[ 6\frac{\ddot{a}}{a} + 3\frac{\dot{a}^2}{a^2} - \frac{\ddot{a}^2}{\dot{a}^2} \Bigr] \Bigr) \phi^2 -\frac{\lambda}{2}\phi^4 \Biggr)
\end{split}
\end{equation}
where we used \Ref{RWpot}, setting $K=0$ for simplicity (we dropped total derivative terms; $\dot a\equiv da/dt$ etc). Note that, in the standard case $\chi=0$, the Lagrangian in \Ref{L} is of the following usual type in dynamical systems: $L=L(a,\dot a,\phi,\dot\phi)$, corresponding to second order differential equations familiar from mechanics. However, in the case $\chi=1$, the Lagrangian also depends on $\ddot a$, and this leads to a 4-th order time evolution equation for $a$; see e.g.\ \cite{Simon} for some general background on such higher-derivative Lagrangian systems. We note that, even though the results described below are well-known in the case $\chi=0$, our method seems different from the ones used in the literature.
By straightforward computations we obtain from \Ref{L} the following Euler-Lagrange equations,
\begin{equation}
\label{ELphi}
\ddot \phi +3H\dot \phi +\Biggl(m^2 +\chi\Bigl[\dot H + 2H^2 -\frac{\dot H^2}{4 H^2}\Bigr]\Biggr)\phi+\lambda\phi^3=0
\end{equation}
and
\begin{equation}
\label{ELa}
\begin{split}
&2\dot H + 3H^2 -\Lambda + \frac1{M_{Pl}^{2}}\Bigl(\frac{\dot\phi^2}2-\frac{m^2}2\phi^2-\frac{\lambda}4\phi^4\Bigr) \\ &+\frac{\chi}{M_{Pl}^{2}}\Biggl( \phi^2\Bigl[ \frac{\ddot H}{2H}+\frac{\dot H^3}{4H^4}-\frac{H^2}2-\frac{\dot H\ddot H}{3H^3} + \frac{5\dot H}{12} -\frac{3\dot H^2}{8H^2}+\frac{\dddot H}{12 H^2}\Bigr]\\ &+\phi\dot\phi\Bigl[\frac{\dot H}{H}-\frac{2H}3 +\frac{\ddot H}{3H^2}-\frac{\dot H^2}{2H^3} \Bigr] +[\phi\ddot\phi+\dot\phi^2]\Bigl[\frac{\dot H}{6H^2}-\frac13\Bigr]\Biggr)=0
\end{split}
\end{equation}
with $H\equiv \dot a/a$. It is interesting to note that only $H$ and its derivatives appear in these equations, and thus the order of the Euler-Lagrange equation corresponding to $a$ is effectively reduced by one. It is important to note that the Lagrangian in \Ref{L} is invariant under time translations, and therefore the time evolution equations in \Ref{ELphi} and \Ref{ELa} allow for a conservation law given by \cite{Simon}
\begin{equation}
\label{Idef}
I \equiv \dot a\left( \frac{\partial L}{\partial \dot a} -\frac{d}{dt}\left(\frac{\partial L}{\partial \ddot a}\right)\right) + \ddot a\frac{\partial L}{\partial \ddot a} + \dot\phi \frac{\partial L}{\partial \dot \phi} -L
\end{equation}
(using the Lagrange equations one easily checks that $dI/dt=0$). By straightforward computations we find that, up to a factor $a(t)^3$, this conservation law can be expressed in terms of $H$ and its derivatives as follows:
\begin{equation}
\label{C}
\begin{split}
\mathcal{I} \equiv -\frac{I}{3 a^3M_{Pl}^2}=& H^2 -\frac{\Lambda}{3}-\frac{1}{6M_{Pl}^2}\Bigl(\dot\phi^2 + m^2\phi^2 + \frac{\lambda}{2}\phi^4 \Bigr)\\ &+\frac{\chi}{M_{Pl}^{2}}\Bigl( \phi\dot\phi \Bigl[\frac16\frac{\dot H}{H}-\frac13H\Bigr]+\phi^2\Bigl[-\frac18\frac{\dot H^2}{H^2}+\frac1{12}\frac{\ddot H}{H} -\frac16H^2 +\frac14\dot H\Bigr] \Bigr) .
\end{split}
\end{equation}
In the standard case $\chi=0$, the two Friedmann equations are equivalent to \Ref{ELa} and $\mathcal{I}=0$: the dynamical system defined by the Lagrangian in \Ref{L} allows for many more solutions than Einstein's equations obtained by varying the action in \Ref{S}, but the additional solutions are eliminated by imposing that the constraint that the value of the conservation law in \Ref{Idef} is zero. {\em We assume that this is true also in the extended case $\chi=1$.} We emphasis that this is a hypothesis but, as we believe, a plausible one. To check it one should generalize the method described in the previous paragraph to the case $\chi=1$, but this we leave for future work.
We thus propose the equations in \Ref{ELphi}, \Ref{ELa}, and $\mathcal{I}=0$ with $\mathcal{I}$ in \Ref{C} and for $\chi=1$, as extension of a standard cosmological model by extrinsic curvature effects. We are mainly interested in this model short after the big bang. We thus make the following ansatz for solutions,
\begin{equation}
\label{ansatz}
\begin{split}
\phi(t) &= t^{-y}\Bigl( \phi_0 + \phi_1 t + \phi_2 t^2 + \phi_3 t^3+\cdots\Bigr) \\
H(t) &= \frac{x}t + H_0 + H_1t + H_2t^2 + H_3t^3+\cdots .
\end{split}
\end{equation}
We call solutions with $x=0$ {\em generic}, and the others {\em scaling solutions}. The reason for these names is as follows: since the equation of motions are invariant under time translations, one can replace in \Ref{ansatz} $t$ by $t-t_{0}$, $t_{0}$ arbitrary, and get another solution. Generically, the fields $a(t)$ and $\phi(t)$ at their derivatives at some time $t_0$ are finite, and the generic solution gives a series representing these fields in the vicinity of $t=t_0$ in terms of appropriate initial conditions at $t=t_0$. For scaling solutions, on the other hand, the time $t=0$ (say) is special: it is the time when the scaling factor vanishes like $(t/t_*)^x$ for some $t_*>0$, and, if $y>0$, the boson field diverges.
\subsection{Solutions}
\label{RWsolutions}
We now discuss solutions of \Ref{ELphi}, \Ref{ELa}, and $\mathcal{I}=0$ with $\mathcal{I}$ in \Ref{C}, for the two cases $\chi=0$ and $\chi=1$ (some details on how we obtained these solutions are given in Appendix~\ref{appRW1}).
For $\chi=0$, we found the following generic solution,
\begin{equation}
\label{generic, chi=0}
\begin{split}
\phi(t) &= \phi_0+\phi_1t-\frac{1}{2}(3H_0\phi_1+[m^2+\lambda \phi_0^2]\phi_0 )t^2+O(t^3)\\
H(t)&=H_0-\frac{\phi_1^2}{2M_{Pl}^2}t+\frac{\phi_1}{2M_{Pl}^2}(3H_0\phi_1+[m^2+\lambda \phi_0^2]\phi_0 )t^2+O(t^3) \\
H_0&=\sqrt{\Lambda/3 +\bigl( \phi_1^2/2 + m^2\phi_0^2/2 + \lambda\phi_0^4/4 \bigr)/(3M_{Pl}^2)}
\end{split}
\end{equation}
(the coefficients of the $O(t^3)$-terms are given in Appendix~\ref{appRW1}, \Ref{O3, chi=0}). This solution depends on two free parameters $\phi_0\equiv \phi(0)$ and $\phi_1\equiv \dot \phi(0)$, and this is the maximum possible number of free parameters: for $\chi=0$, \Ref{ELphi} and \Ref{ELa} are second- and first order differential equations, allowing for two- and one integration constants, respectively, and the constraint $\mathcal{I}=0$ reduces the number of free parameters by one.
In the standard case $\chi=0$, we did not find any solutions as in \Ref{ansatz} with $x\neq 0$. However, in the extended case $\chi=1$, we found two such solutions: one with $x=3/4$, $y=1$ given by
\begin{equation}
\label{34}
\begin{split}
\phi(t)=& \sqrt{\frac2{\lambda}}\Bigl( \frac1{4t} - \Bigl[\frac{2 m^2}{15}-\frac{126\lambdaM_{Pl}^2}{65} \Bigr] t + O(t^3) \Bigr) \\
H(t)=& \frac{3}{4t} + \Bigl( \frac{m^2}{5}-\frac{594\lambdaM_{Pl}^2}{65}\Bigr)t + O(t^3),
\end{split}
\end{equation}
and another one with $x=(1+\sqrt{3})/4=0.683(0)$, $y=0$ given by\footnote{We write $0.683(0)$ for a numerical value $0.6830\pm 0.0001$.}
\begin{equation}
\label{xc}
\begin{split}
\phi(t) =& M_{Pl}\Bigl(1.50(2) - \Bigl[0.228(0)\,{m}^{2} + 0.402(0)\,\lambdaM_{Pl}^2 + 0.0883(5)\,{\Lambda}\Bigr]t^2 + O(t^4) \Bigr) \\
H(t) = &\frac{0.683(0)}{t} - \Bigl( 0.0142(5)\,{m}^{2}+ 0.120(0)\lambdaM_{Pl}^2 - 0.0690(4)\,\Lambda \Bigr)t + O(t^3) .
\end{split}
\end{equation}
Note that the solution in \Ref{34} exists only if $\lambda>0$, whereas the solution in \Ref{xc} exists even if $\Lambda=\lambda=0$. Thus the latter solution is more robust and, as we believe, more interesting. We therefore give formulas for the coefficients of the $O(t^3)$-terms in \Ref{xc} in Appendix~\ref{appRW1}, \Ref{xcO3}.
For $\chi=1$ we found the following generic solution,
\begin{equation}
\label{generic, chi=1}
\begin{split}
\phi(t)= & \phi_0+\phi_1t -\frac12\Bigl(3H_0\phi_1+\phi_0\Bigl[m^2+\lambda \phi_0^2 +H_1 + 2H_0^2 -\frac{H_1^2}{4H_0^2} \Bigr] \Bigr)t^2 + O(t^3) \\
H(t)= & H_0+H_1t+\Bigl( \Bigl[m^2 +\frac{\lambda}2\phi_0^2+H_0^2-\frac32 H_1\Bigr]H_0 +\frac{3H_1^2}{4H_0}\\ & +(2H_0^2-H_1)\frac{\phi_1}{\phi_0}+\bigl(\phi_1^2+2\LambdaM_{Pl}^2-6H_0^2M_{Pl}^2 \bigr)\frac{H_0}{\phi_0^2}\Bigr) t^2+O(t^3)
\end{split}
\end{equation}
which depends on four free parameters: $\phi_0\neq 0$, $\phi_1$ as before and, in addition, $H_0\equiv H(0)$ and $H_1\equiv \dot H(0)$. Note that this is the maximum number of free parameters (since \Ref{ELa} now is third order, we have two more free parameters).
We note the {\em static} solutions, i.e., $\phi(t)=\phi_0$ and $H(t)=H_0$ independent of $t$, which are interesting special cases of the generic solutions above and exist if $m^2<0$, $\lambda>0$ ("Mexican hat potential") and $\Lambda>m^4/(4\lambdaM_{Pl}^2)$. They are given by\footnote{Note that for every solution $\phi(t)$, $H(t)$ we give, $-\phi(t)$, $H(t)$ is also a solution.}
\begin{equation}
\label{chi=0, static}
\phi_0 = \sqrt{-\frac{m^2}{\lambda}},\quad H_0=\sqrt{\frac{4\lambda\Lambda-m^4/M_{Pl}^2}{12\lambda}} \qquad\quad (\chi=0)
\end{equation}
and
\begin{equation}
\label{chi=1, static}
\phi_0 = \sqrt{\frac{2(-3m^2-2\Lambda)}{6\lambda +m^2/M_{Pl}^2}},\quad H_0=\sqrt{\frac{4\lambda\Lambda-m^4/M_{Pl}^2}{2(6\lambda +m^2/M_{Pl}^2)}} \qquad\quad (\chi=1)
\end{equation}
in the standard- and extended cases, respectively. Note that, for $\chi=1$, we have the additional conditions $\Lambda<-3m^2/3$ and $\lambda>-m^2/(6M_{Pl}^2)$ for these solutions to exist. Moreover, for $\Lambda>0$ and in both cases $\chi=0$ and $\chi=1$, there exists also the static solution with $\phi_0=0$ and $H_0=\sqrt{\Lambda/3}$.
\subsection{Discussion}
\label{RWdiscussion}
As already mentioned, the generic solutions allows us to compute $a(t)$ and $\phi(t)$ in some interval around an arbitrary initial time $t=t_0$ from its initial conditions at $t=t_0$. In the standard case, they allow us to understand the importance of the so-called "slow roll" condition for inflation in this model \cite{LythLiddle09} in the following way: the equations of motion of the boson field $\phi(t)$ are identical with Newton's equations for a particle moving in one dimension under the influence of a friction term and a potential $\mathcal{V}(\phi)=m^2\phi^2/2+\lambda\phi^4/4$: $\ddot\phi(t)=-3H(t)\dot\phi(t)-\mathcal{V}'(\phi(t))$, and \Ref{generic, chi=0} implies $\dot H(t)=-\dot\phi(t)^2/(2M_{Pl}^2)$ and $\ddot H(t)=-6H(t)\dot H(t)+d{\mathcal{V}(\phi(t))}/dt$ (these equations can also be obtained directly from the Friedmann equations). This suggests that, starting from arbitrary initial conditions $\phi(0)$ and $\dot\phi(0)$ at some time $t=0$, $\phi(t)$ will increase with decreasing $\dot \phi(t)$ until $\dot\phi(t_1)=0$ at some time $t=t_1$, and, at this time, $\dot H(t_1)=\ddot H(t_1)=0$. In the vicinity of $t=t_1$,\footnote{We checked that the following result holds true for arbitrary differentiable boson potentials $\mathcal{V}(\phi)$.}
\begin{equation}
\label{series}
\begin{split}
\phi(t)&=\phi_0 -\frac12\mathcal{V}'(\phi_0)(t-t_1)^2+O((t-t_1)^2)\\
\ln\Bigl( \frac{a(t)}{a(t_1)} \Bigr) &=H_0(t-t_1) - \frac{\mathcal{V}'(\phi_0)^2}{24M_{Pl}^2}(t-t_1)^4 + O((t-t_1)^5)
\end{split}
\end{equation}
with $\phi_0\equiv \phi(t_1)$ and $H_0=\sqrt{\Lambda/3+\mathcal{V}(\phi_0)/(3M_{Pl}^2)}$, i.e., the time evolution of the scaling factor can be well approximated by the exponential law $a(t)\approx a(t_1)\exp\bigl( H(t_1)(t-t_1)\bigr)$, up to corrections that are negligible as long as $\dot\phi(t)^2\ll 24M_{Pl}^2H(t_1)/|t-t_1|$. We computed and plotted the approximation of the solution obtained by extending this series in \Ref{series} to higher orders and ignoring the $O((t-t_1)^n)$-terms, for different values of $n$, up to $n=8$. The result suggests that these series have a finite radius of convergence which, however, is often smaller than the full time interval where inflation occurs.
We now discuss the extended case $\chi=1$ and how it differs from the standard case $\chi=0$. Obviously, there are {\em two} essential differences: firstly, the generic solution in \Ref{generic, chi=1} depends on {\em four} (rather than two for $\chi=0$) initial conditions, and secondly, there exist scaling solutions (which do not exist for $\chi=0$). In the next two paragraph we discuss these two differences and possible implications.
The generic solution in \Ref{generic, chi=1} depending on four initial conditions is a consequence of the the Lagrangian depending also on $\ddot a(t)$ As illustrated in a simple toy model in Appendix~\ref{appToyModel}, even if such higher-order term in the Lagrangian is very small, it can lead to a much richer {\em qualitative} behavior of the system. Thus, at first sight, it seems that the model with $\chi=1$ is less predictive than the standard model with $\chi=0$. However, our toy model also suggests that it is possible to restrict the initial conditions of the model with $\chi=1$ so as to obtain solutions that are similar to the ones for $\chi=0$. One natural way to impose such a restriction is suggested by the generic solution in \Ref{generic, chi=1}: obviously, this solution is not well-defined if $\phi_0=0$. Using translation invariance in time, we conclude that a generic solution such that $\phi(t_0)=0$ at some time $t=t_0$ can exist only if we impose some restrictions of the free parameters. One can check that these restrictions on initial conditions at time $t=t_0$ are
\begin{equation}
\label{H01}
H_0^2=\frac{\Lambda}{3}+\frac{\phi_1^2}{6M_{Pl}^2},\quad H_1=-\frac{2\phi_1^2(\phi_1^2+2\LambdaM_{Pl}^2)}{3M_{Pl}^2(3\phi_1^2+4\LambdaM_{Pl}^2)}
\end{equation}
with $H_0\equiv H(t_0)$ etc.\ (see Appendix~\ref{appRW1} for derivation). This leads to
\begin{equation}
\label{phi0=0}
\begin{split}
\phi(t)=\phi_1 (t-t_0) -\frac{3}{2}\phi_1H_0(t-t_0)^2 + O((t-t_0)^3) , \quad
H(t) = H_0 + H_1 (t-t_0) \\ + \frac{4H_0^3 \phi_1^2(19\phi_1^4+36\phi_1^2M_{Pl}^2\Lambda+32(\LambdaM_{Pl}^2)^2)}{(3\phi_1^2+4\LambdaM_{Pl}^2)^2(5\phi_1^2+4\LambdaM_{Pl}^2))}(t-t_0)^2+O((t-t_0)^3)
\end{split}
\end{equation}
with $H_0$ and $H_1$ as in \Ref{H01}. We thus propose the following hypothesis: {\em Only solutions that remain well-defined when $\phi(t)\to 0$ as $t\to t_0$, for some time $t_0$, can describe the evolution of the universe}. This reduces the number of free parameters to two, which is equal to the number of free parameters in the standard case $\chi=0$: the time $t=t_0$ where the boson field $\phi(t)$ vanishes, and the time derivative $\phi_1\equiv \dot\phi(t_0)$ of the boson field at that time. We believe that it is these restricted solutions of the extended case $\chi=1$ that behave, at larger times, similar to the ones on the standard case $\chi=0$. We note that there might be also solutions where $\phi(t)>0$ for all times, but these we do not know how to restrict in a natural way.
We believe that the existence of scaling solutions in the extended case $\chi=1$, which do not have any analogue for $\chi=0$, could make the former model more predictive than the latter in the following way: for non-linear dynamical systems, scaling solutions often are attractors that capture the qualitative behavior of a large class of solutions (see e.g.\ \cite{Barenblatt}). As mentioned already above, the scaling solution in \Ref{xc} is more robust and thus, as we believe, more interesting than the one in \Ref{34}. We thus believe that it would be interesting to explore the validity of the following hypothesis: {\em Solutions that can describe the universe immediately after the big bang approach the scaling solution in \Ref{xc} as $t\to 0^+$.} Combining this hypothesis with the one above we get the following picture: at some time $t_0>0$, $\phi(t_0)=0$, and in some interval the time evolution is described by \Ref{H01}--\Ref{phi0=0}. The free parameters $\phi_1$ and $t_0$ of this solution are then to be fixed such that, as $t\to 0^+$, this solution approaches the one in \Ref{xc}. For $t>t_0$, the boson field $\phi(t)$ grows with decreasing $\dot\phi(t)$ until $\dot\phi(t)=0$ at some time $t=t_1$, and close to this time the solution is, as we expect, similar to the one in \Ref{series}. If or not this picture holds true can be determined by a numerical solution, but this we leave to future work. We expect that the solution close to the time $t=t_1$ describes inflation, similarly as discussed after \Ref{series}. One important question is if such a solution can describe inflation (solving the horizon problem etc.\ \cite{LythLiddle09}) even in some time interval short after $t=0$ or not, i.e., if such a solution has one or two periods of inflation. At first sight it seems the answer to this question is negative: $a(t)\to (t/t_*)^x$ as $t\to 0^+$ with $x=0.683(0)$, and previous work on power-law inflation \cite{AW,LM} suggests that inflation requires $x>1$. However, in our scaling solution, $a(t)$ is well approximated by $(t/t_*)^x$ only in a small time interval, and for larger times the behavior is more complicated. We thus believe that this question is open.
\section{Conclusions}
\label{sec7}
In this paper we proposed that the induced potential, which is well-established in constrained quantum mechanics, could also be relevant in brane-world scenarios where space-time is embedded in a higher dimensional ambient space. We showed that, by assuming that the propagating degrees of freedom are restricted to space-time by a strong confining potential, one finds that the Klein-Gordon equation on space-time is indeed modified by a induced potential term which does not depend of the details of the confining potential. While the formula for the induced potential we obtain is the same as in constrained quantum mechanics, the physical arguments to derive this potential are different.
One striking feature of the induced potential is that it does not only depend on intrinsic- but also on extrinsic geometric properties of the embedded space-time. Thus, if the induced potential has observable consequences, it offers the intriguing possibility to test brane-world scenarios experimentally. As examples we computed the induced potential for Schwarzschild- and Robertson-Walker space-times. Our results suggest that, while the induced potential is usually negligibly small, it might be relevant in extreme situations like the early universe or in regions close to primordial black holes. We also proposed and studied a model for cosmological inflation with the effect of the induced potential included. At first sight this generalized model seems to be less predictive (since the generic solution of the equations of motion depend of four initial conditions, rather than two in the standard case). However, we found a natural condition reducing the number of initial conditions to two. Moreover, the extended model is different from the standard one also in that it allows for scaling solutions that do not depend on any free parameter. These scaling solution describe an expanding universe with a size approaching zero at some initial time. Our results suggest that, in the extended model, one only has to postulate a big bang, and the time evolutions of the universe is fixed, without much freedom to vary initial conditions. Anyway, it seems worthwhile to study this model further.
For simplicity we restricted ourselves in this paper to the case where the ambient space is flat. It would be interesting to generalize our results in Section~\ref{sec2} to cases where the ambient space is curved. We expect that the induced potential thus obtained is the obvious generalization of one known in constrained quantum mechanics; see e.g.\ \cite{Mitchell}, Equation~(3.37).
|
\section{Introduction}
\label{sec:intro}
The importance of the representation theory of the three-dimensional
rotation group \cite{edm96,wig59,jud75,bie85a,bie85b,var88,ham89} in
the study of all natural quantum systems hardly needs to be mentioned.
That central role is due to the fact that the four fundamental
interactions of nature, and their effective interactions relevant to
nuclear, atomic and molecular physics, are all invariant under the
rotation group. It is also for that reason that partial-wave
expansions are essential tools in the analysis of both classical and
quantum scattering processes.
The algebraic aspects of the computation of partial-wave expansions
with spin states are best codified by the addition theorems for spin
spherical harmonics. Spin-$s$ spherical harmonics $Y^{\ell
s}_{jj_z}(\verr{})$ \cite{bie85a,var88} are the angular-momentum
eigenfunctions relevant to the description of spin-$s$ particles
subject to spin-dependent central interactions.\footnote{Familiar
examples of spin-$s$ spherical harmonics are ordinary scalar
spherical harmonics, $Y^{\ell 0}_{\ell\ell_z}(\verr{}) =
Y_{\ell\ell_z}(\verr{})$, and vector spherical harmonics
\cite{bie85a,var88,gal90,jackxx}, $Y^{\ell 1}_{j j_z}(\verr{})$.}
As such, they are of interest in their own right beyond their
above-mentioned specific application to partial-wave expansions that
constitutes our main motivation for their study. In this paper,
together with its second part \cite{bou10} (hereafter referred to as
II), we develop a systematic framework to derive addition theorems for
spin spherical harmonics, and employ it to obtain those addition
theorems for low spins, $0<s', s\leq 3/2$, and for arbitrary integer
spin $s'$ if $s=0$.
In this first part we establish the preliminary results that will be
used as basic building blocks for constructing addition theorems for
spin spherical harmonics in II. We consider first the factorization
of orbital and spin degrees of freedom in products of Clebsch-Gordan
(henceforth CG) coefficients of the form
$\CG{\ell'}{\ell'_z}{s'}{s'_z}{j}{j_z}
\CG{\ell}{\ell_z}{s}{s_z}{j}{j_z}$. We obtain a general factorization
result, make its tensor and spinor structure explicit, and discuss the
particular cases relevant to spins 1/2, 1 and 3/2. Such products of
CG coefficients occur in the computation of physical quantities, such
as scattering amplitudes or other matrix elements, related to
transitions among states with definite orbital and spin quantum
numbers. Thus, these results may be of interest independently of the
addition theorems considered here and in II. We study also the matrix
elements of arbitrary tensor products of orbital operators between
angular-momentum projections of position eigenstates, for which we
obtain completely general results. These are the other main
ingredient in our derivation of addition theorems for spin spherical
harmonics in II. Those matrix elements are closely related to bilocal
spherical harmonics and other bilocal sums of ordinary spherical
harmonics, as discussed in detail in II, so we expect that these
results and the approach used to derive them should be applicable in
other contexts as well.
The outline of the paper is as follows. In the next section we
discuss the factorization of orbital and spin degrees of freedom in
products of CG coefficients. In section \ref{sec:proj} we
introduce the angular-momentum projector operator and obtain general
expressions for its matrix elements with orbital tensor operators. In
appendix \ref{sec:notatio} we state our notation and conventions. In
appendix \ref{sec:standard} we give the definitions and main
properties of the standard bases of irreducible tensors and spinors,
of arbitrary spin, used as spin wavefunctions throughout the paper. A
detailed list of general results for reduced matrix elements of
arbitrary tensor products of orbital and spin operators, needed in the
main sections of the paper, are given in appendix \ref{sec:matele}.
Finally, appendix \ref{sec:appa} gathers some ancillary calculations
needed in sect.\ \ref{sec:proj}.
\section{Factorization of orbital and spin dependence}
\label{sec:fac}
We begin by considering products of CG coefficients of the form
\begin{equation}
\label{eq:additionintro}
\begin{aligned}
S(j,\ell',\ell,s',s;\ell'_z,\ell_z,s'_z,s_z) &= \sum_{j_z=-j}^j
\CG{\ell'}{\ell'_z}{s'}{s'_z}{j}{j_z}
\CG{\ell}{\ell_z}{s}{s_z}{j}{j_z} \\
&=
\CG{\ell'}{\ell'_z}{s'}{s'_z}{j}{\ell'_z+s'_z}
\CG{\ell}{\ell_z}{s}{s_z}{j}{\ell_z+s_z}
\delta_{\ell'_z+s'_z,\ell_z+s_z}~,
\end{aligned}
\end{equation}
with $\ell'$, $\ell$ integer and $s'$, $s$ integer or half-integer.
We will usually write $S(j,\ell',\ell,s',s)$, omitting the last
arguments for brevity. For the purpose of obtaining addition theorems
for spin spherical harmonics we need to rewrite $S(j,\ell',\ell,s',s)$
as a sum of terms with a completely factorized dependence on orbital
and spin angular-momentum projections, and to make explicit the
tensor structure of those terms. The first goal is achieved by making
use of relation (\ref{eq:6j-1}) to write (see appendix
\ref{sec:notatio} for our notation and conventions),
\begin{equation}
\label{eq:6j-2}
\begin{gathered}
\begin{aligned}
S(j,\ell',\ell,s',s) &=
(-1)^{s'_z+s_z}(-1)^{\ell+s'-j}
\frac{2j+1}{\sqrt{(2\ell'+1)(2s'+1)}}
\sum_{\Delta=\Delta_\mathrm{min}}^{\Delta_\mathrm{max}} (2\Delta+1)
\SJ{\ell'}{\Delta}{\ell}{s}{j}{s'}
\\
&\quad\times
\CG{\ell}{\ell_z}{\Delta}{\Delta\ell_z}{\ell'}{\ell'_z}
\CG{s}{s_z}{\Delta}{\Delta s_z}{s'}{s'_z}~,
\end{aligned}\\
\Delta_\mathrm{min} = \max\{|\Delta\ell|,|\Delta s|\}~,
\qquad
\Delta_\mathrm{max} = \min\{\ell'+\ell,s'+s\}~,
\qquad
\Delta\ell_z + \Delta s_z =0~.
\end{gathered}
\end{equation}
In order to make the tensor structure of each term on the r.h.s.\
explicit, we use the Wigner-Eckart (henceforth WE) theorem
\cite{edm96,wig59,jud75,bie85a,bie85b,var88,gal90} to write
the CG coefficients in terms of appropriate irreducible tensor
operators,
\begin{equation}
\label{eq:CGmaster}
\begin{gathered}
\begin{aligned}
S(j,\ell',\ell,s',s) &=
\sum_{\Delta=\Delta_\mathrm{min}}^{\Delta_\mathrm{max}}
C^{s's\Delta}_{\ell'\ell j} \langle\ell',\ell'_z | \veps{i_1\ldots
i_\Delta}(\Delta\ell_z) \verr{i_1}\ldots \verr{i_{|\Delta\ell|}}
L^{i_{|\Delta\ell|+1}} \ldots L^{i_\Delta} | \ell, \ell_z\rangle\\
&\quad\times
\langle s',s'_z | \veps{h_1\ldots
h_\Delta}(\Delta\ell_z)^* T^{h_1}\ldots T^{h_{|\Delta s|}}
S^{h_{|\Delta s|+1}} \ldots S^{h_\Delta} | s, s_z\rangle ~,
\end{aligned}\\
\begin{aligned}
C^{s's\Delta}_{\ell'\ell j} &= (-1)^{2s'_z}(-1)^{\ell+s'-j}
\frac{2j+1}{\sqrt{(2\ell'+1) (2s'+1)}} (2\Delta+1)
\SJ{\ell'}{\Delta}{\ell}{s}{j}{s'} \\
&\quad\times \left(\langle\ell'|| \veps{j_1\ldots
j_\Delta} \verr{j_1}\ldots \verr{j_{|\Delta\ell|}}
L^{j_{|\Delta\ell|+1}} \ldots L^{j_\Delta} || \ell\rangle
\langle s'|| \veps{k_1\ldots
k_\Delta} T^{k_1}\ldots T^{k_{|\Delta s|}}
S^{k_{|\Delta s|+1}} \ldots S^{k_\Delta} || s\rangle
\right)^{-1}~,
\end{aligned}
\end{gathered}
\end{equation}
where $\Delta_{\mathrm{min},\mathrm{max}}$ are as in (\ref{eq:6j-2}),
and our notation for the standard basis tensors $\veps{i_1\ldots i_n}$
and for irreducible tensor operators is explained in appendix
\ref{sec:standard}. The operators appearing in (\ref{eq:CGmaster})
are the orbital $\vec{L}$ and spin $\vec{S}$ angular momenta, the
position versor $\verr{}=\vec{r}/r$, and the spin-transition operator
$\vec{T}$ defined here as the spin analog of $\verr{}$ (see sect.\
\ref{sec:spintrans} below for a detailed discussion of this operator).
The reduced matrix elements appearing in (\ref{eq:CGmaster}) are
explicitly given for arbitrary values of their parameters in appendix
\ref{sec:matele}.
Whereas (\ref{eq:CGmaster}) is completely general, it is not yet
explicit enough for the purpose of deriving addition theorems for spin
spherical harmonics. From a practical point of view, we need only
consider low values of $s'$, $s$ since quantum states with high values
of spin occur infrequently in nature. In this paper we restrict
ourselves to $s'$, $s=0$, 1/2, 1, 3/2, and $|\Delta s|\leq 1$. For
those spins the coefficients $C^{s's\Delta}_{\ell'\ell j}$ can be
drastically simplified by exploiting the fact that $j-\ell'$ and
$j-\ell$ can take only a small set of values. The spin matrix
elements in (\ref{eq:CGmaster}) can be compactly evaluated in terms of
standard tensors and spinors, and spin matrices. (The orbital matrix
elements could, in principle, also be expressed in that way, but it
would be impractical because $\ell'$, $\ell$ are not bounded above
thus requiring tensor wave functions of arbitrarily large rank. We
evaluate those matrix elements with a completely different technique
below in section \ref{sec:proj}.) Clearly, we must discuss separately
the different values of $s'$ and $s$, as those will determine the kind
of spin spherical harmonics involved in the addition theorem. For
fixed $s'$ and $s$, different values of $\Delta\ell$ correspond to
different addition theorems, so we must consider those cases
separately as well. Furthermore, both in addition theorems for spin
spherical harmonics and in partial wave expansions, terms with
different $|\Delta\ell|$ have different tensor and spin structure.
\subsection{$\boldsymbol{s'=1/2=s}$}
\label{sec:1half}
For $s'=1/2=s$ there are two possible values of $|\Delta\ell|=0$, 1.
\paragraph*{$\boldsymbol{\Delta\ell =0}$}
\label{sec:1half-0}
In the case $s=1/2$, $j$ can take only two values and
(\ref{eq:CGmaster}) reduces to
\begin{equation}
\label{eq:1half-1}
\begin{gathered}
S(j,\ell,\ell,1/2,1/2) = C^{\frac{1}{2}\frac{1}{2}0}_{\ell\ell j} \delta_{\ell'_z \ell_z}
\delta_{s'_z s_z} + C^{\frac{1}{2}\frac{1}{2}1}_{\ell\ell j} \langle \ell,\ell'_z |
\veps{k}(\Delta\ell_z) L^k | \ell,\ell_z \rangle\, \veps{h}(\Delta\ell_z)^*
\vchi{A}(s'_z)^* \frac{1}{2}\sigma^h_{AB}\vchi{B}(s_z) ~,\\
C^{\frac{1}{2}\frac{1}{2}0}_{\ell\ell j} = \frac{j+1/2}{2\ell+1}~,
\qquad
C^{\frac{1}{2}\frac{1}{2}1}_{\ell\ell j} =
\frac{2(j-\ell)}{\ell+1/2}~,
\end{gathered}
\end{equation}
which is a well-known result (see e.g., \cite{nac90}). Notice that,
due to the conservation of $j_z$, the product
$\veps{k}(\Delta\ell_z)\veps{h}(\Delta\ell_z)^*$ in (\ref{eq:1half-1}) can be replaced
by $\delta^{kh}$. We quote here also the simplified form of
(\ref{eq:6j-2}) in this case,
\begin{equation}
\label{eq:1half-2}
S(j,\ell,\ell,1/2,1/2) = C^{\frac{1}{2}\frac{1}{2}0}_{\ell\ell j} \delta_{\ell'_z \ell_z}
\delta_{s'_z s_z} + \frac{\sqrt{3}}{2} \sqrt{\ell(\ell+1)} C^{\frac{1}{2}\frac{1}{2}1}_{\ell\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1/2}{s'_z}{1}{\Delta\ell_z}{1/2}{s_z}~,
\end{equation}
which results from (\ref{eq:1half-1}) by applying the WE theorem.
\paragraph*{$\boldsymbol{|\Delta\ell|=1}$}
\label{sec:1half-1}
In this case $S(j,\ell',\ell,1/2,1/2)$ can be non-vanishing only if
$\ell'=\ell+1$ and $j=\ell+1/2$, or if $\ell'=\ell-1$ and
$j=\ell-1/2$. Equation (\ref{eq:CGmaster}) takes the simplified form,
\begin{equation}
\label{eq:1half-3}
S(j,\ell_1,\ell,1/2,1/2) = C^{\frac{1}{2}\frac{1}{2}1}_{\ell_1\ell j}
\langle \ell_1,\ell'_z | \veps{k}(\Delta\ell) \verr{k} | \ell,\ell_z \rangle\,
\veps{h}(\Delta\ell)^* \vchi{A}(s'_z)^*
\frac{1}{2}\sigma^h_{AB}\vchi{B}(s_z) ~,
\quad
C^{\frac{1}{2}\frac{1}{2}1}_{\ell_1\ell j} = -2~,
\end{equation}
which is appropriate to derive an addition theorem for spin-1/2
spherical harmonics. As in the previous case, the product of standard
versors can be substituted by $\delta^{kh}$. Similarly,
(\ref{eq:6j-2}) reduces to
\begin{equation}
\label{eq:1half-4}
S(j,\ell_1,\ell,1/2,1/2) = -\Delta\ell\sqrt{\frac{3}{2}}
\sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1/2}{s'_z}{1}{\Delta\ell_z}{1/2}{s_z} ~.
\end{equation}
As a side remark, we notice that in this case $(2\langle\ell\rangle+1)/(2\ell_1+1)
= (2j+1)/(2j+1+\Delta\ell)$, which provides an alternate form for
(\ref{eq:1half-4}).
\subsection{$\boldsymbol{s'=1=s}$}
\label{sec:1}
The general result (\ref{eq:CGmaster}) is greatly simplified in the
case of $S(j,\ell',\ell,1,1)$ because, for fixed $\ell$, $j-\ell$ can
take only the three values $\pm1$, 0. The spin matrix elements are
given by
\begin{equation}
\label{eq:1-1}
\begin{aligned}
\langle 1,s'_z | \veps{h}(\Delta\ell_z)^* S^h | 1,s_z\rangle = -i
\veps{h}(\Delta\ell_z)^* \left(\veps{}(s'_z)^* \wedge
\veps{}(s_z)\right)^h~, \\
\langle 1,s'_z | \veps{h_1h_1}(\Delta\ell_z)^* S^{h_1}S^{h_2} |
1,s_z\rangle = -\veps{h_1h_2}(\Delta\ell_z)^* \veps{h_1}(s'_z)^*
\veps{h_2}(s_z)~.
\end{aligned}
\end{equation}
There are three possible values for $\Delta\ell=0$, 1, 2.
\paragraph*{$\boldsymbol{|\Delta\ell|=0}$}
\label{sec:1-1}
In this case (\ref{eq:CGmaster}) can be written as
\begin{equation}
\label{eq:1-2}
\begin{gathered}
\begin{aligned}
S(j,\ell,\ell,1,1) &= C^{110}_{\ell\ell j} \delta_{\ell'_z\ell_z}
\delta_{s'_zs_z} - C^{111}_{\ell\ell j} \langle \ell,\ell'_z |
\veps{k}(\Delta\ell_z) L^k | \ell,\ell_z\rangle\,\veps{h}(\Delta\ell_z)^*
i \left(\veps{}(s'_z)^* \wedge \veps{}(s_z)\right)^h\\
&\quad - C^{112}_{\ell\ell j} \langle \ell,\ell'_z |
\veps{k_1k_2}(\Delta\ell_z) L^{k_1} L^{k_2} | \ell,\ell_z\rangle\,\veps{h_1h_2}(\Delta\ell_z)^*
\veps{h_1}(s'_z)^* \veps{h_2}(s_z)~,
\end{aligned}\\
C^{110}_{\ell\ell j} = \frac{1}{3} \frac{2j+1}{2\ell+1}~,
\quad
C^{111}_{\ell\ell j} = \frac{2}{D_{\ell j}} (2j+1)
((j-\ell)(j+\ell+1)+1)~,
\quad
C^{112}_{\ell\ell j} = \frac{4}{D_{\ell j}} (2j+1)~,\\
D_{\ell j} = (-1)^{j-\ell-1} \left((j-\ell)^2+1\right)(j+\ell+2)
(j+\ell+1) (j+\ell)~.
\end{gathered}
\end{equation}
This form of (\ref{eq:CGmaster}) is needed in the derivation of
addition theorems for vector spherical harmonics. Notice that in the
term multiplied by $C^{111}_{\ell\ell j}$ we can replace
$\veps{k}(\Delta\ell_z) \veps{h}(\Delta\ell_z)^*$ by $\delta^{kh}$, and in the term
multiplied by $C^{112}_{\ell\ell j}$ we can rewrite the matrix
elements as
\begin{equation}
\label{eq:1-3}
\langle \ell,\ell'_z | \veps{k_1k_2}(\Delta\ell_z)L^{k_1}L^{k_2} |
\ell,\ell_z \rangle\, \veps{h_1h_2}(\Delta\ell_z)^*
\veps{h_1}(s'_z)^* \veps{h_2}(s_z) =
\langle \ell,\ell'_z |\frac{1}{2}L^{\{p} L^{q\}_0} |\ell,\ell_z \rangle\,
\veps{p}(s'_z)^* \veps{q}(s_z)~.
\end{equation}
The relation (\ref{eq:6j-2}) in this case reduces to
\begin{equation}
\label{eq:1-4}
\begin{aligned}
S(j,\ell,1) &=
C^{110}_{\ell\ell j}\delta_{\ell'_z\ell_z}\delta_{s'_zs_z}
+\sqrt{2} \sqrt{\ell(\ell+1)}C^{111}_{\ell\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1}{s'_z}{1}{\Delta\ell_z}{1}{s_z} \\
&\quad + \frac{\sqrt{10}}{6} \sqrt{\ell(\ell+1)}
\sqrt{(2\ell-1)(2\ell+3)} C^{112}_{\ell\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1}{s'_z}{2}{\Delta\ell_z}{1}{s_z}~,
\end{aligned}
\end{equation}
with $C^{11\Delta}_{\ell\ell j}$ as in (\ref{eq:1-2}).
\paragraph*{$\boldsymbol{|\Delta\ell|=1}$}
\label{sec:1-2}
In this case $S(j,\ell',\ell,1,1)$ can be non-vanishing only if
$\ell'=\ell+1$ and $j=\ell+1$, $\ell$, or if $\ell'=\ell-1$ and
$j=\ell$, $\ell-1$. With that discrete set of values for $j-\ell$,
the spin matrix elements (\ref{eq:1-1}), and the reduced matrix
elements in appendix \ref{sec:matele}, from (\ref{eq:CGmaster}) we
obtain
\begin{equation}
\label{eq:1-5}
\begin{gathered}
\begin{aligned}
S(j,\ell_1,\ell,1,1) &= -C^{111}_{\ell_1\ell j}
\langle\ell_1,\ell'_z|\veps{k}(\Delta\ell_z)\verr{k}|\ell,\ell_z\rangle\,
i \veps{h}(\Delta\ell_z)^*
(\veps{}(s'_z)^*\wedge\veps{}(s_z))^h\\
&\quad - C^{112}_{\ell_1\ell j}
\langle\ell_1,\ell'_z|\veps{k_1k_2}(\Delta\ell_z)\verr{k_1}L^{k_2}|\ell,\ell_z\rangle
\,
\veps{h_1h_2}(\Delta\ell_z)^* \veps{h_1}(s'_z)^*\veps{h_2}(s_z)~,
\end{aligned}\\
C^{111}_{\ell_1\ell j} = -\frac{1}{2\langle\ell\rangle+1} \sqrt{2j+1}
\sqrt{2j-\langle\ell\rangle+1/2}~,
\quad
C^{112}_{\ell_1\ell j} = -4 \frac{(j-\langle\ell\rangle)}{2\langle\ell\rangle+1}
\sqrt{\frac{2j+1}{2j-\langle\ell\rangle+1/2}} ~.
\end{gathered}
\end{equation}
As before, on the first line of (\ref{eq:1-5}) we can replace
$\veps{k}(\Delta\ell_z) \veps{h}(\Delta\ell_z)^*$ by $\delta^{kh}$, and on the
second line we can rewrite the matrix elements as
\begin{equation}
\label{eq:1-6}
\langle \ell',\ell'_z | \veps{k_1k_2}(\Delta\ell_z)\verr{k_1}L^{k_2} |
\ell,\ell_z \rangle\, \veps{h_1h_2}(\Delta\ell_z)^*
\veps{h_1}(s'_z)^* \veps{h_2}(s_z) =
\langle \ell',\ell'_z |\frac{1}{2}\verr{\{p} L^{q\}_0} |\ell,\ell_z \rangle\,
\veps{p}(s'_z)^* \veps{q}(s_z)~.
\end{equation}
The form of (\ref{eq:6j-2}) in this case is
\begin{equation}
\label{eq:1-7}
\begin{aligned}
S(j,\ell_1,\ell,1,1) &=
\Delta\ell \sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}} \left(
\rule{0pt}{18pt}
C^{111}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1}{s'_z}{1}{\Delta\ell_z}{1}{s_z}
\right.\\
&\quad \left. +\frac{1}{4}\sqrt{\frac{5}{3}}
\sqrt{(2\langle\ell\rangle-1)(2\langle\ell\rangle+3)} C^{112}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1}{s'_z}{2}{\Delta\ell_z}{1}{s_z}\right)~,
\end{aligned}
\end{equation}
with the coefficients (\ref{eq:1-5}).
\paragraph*{$\boldsymbol{|\Delta\ell|=2}$}
\label{sec:1-3}
In this case $S(j,\ell',\ell,1,1)$ can be non-vanishing only if
$\ell'=\ell+2$ and $j=\ell+1$, or if $\ell'=\ell-2$ and $j=\ell-1$~.
Using (\ref{eq:1-1}), (\ref{eq:redmat5}) and (\ref{eq:redmat3}), from
(\ref{eq:CGmaster}) and (\ref{eq:6j-2}) we obtain
\begin{equation}
\label{eq:1-8}
\begin{aligned}
S(j,\ell_2,\ell,1,1) &= -C^{112}_{\ell_2\ell j}
\langle\ell_2,\ell'_z|\veps{k_1k_2}(\Delta\ell_z)\verr{k_1}\verr{k_2}
|\ell,\ell_z\rangle \,
\veps{h_1h_2}(\Delta\ell_z)^* \veps{h_1}(s'_z)^*\veps{h_2}(s_z)\\
&=\sqrt{\frac{5}{3}} \sqrt{\frac{j(j+1)}{(2j+1)(2\ell_2+1)}}
C^{112}_{\ell_2\ell j} \CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_2}{\ell'_z}
\CG{1}{s'_z}{2}{\Delta\ell_z}{1}{s_z}~,\\
C^{112}_{\ell_2\ell j} &= \frac{2j+1}{\sqrt{j(j+1)}}~.
\end{aligned}
\end{equation}
The first line is the form needed for an addition theorem for
vector spherical harmonics. Notice that the matrix element on that
line is traceless, because $\ell'\neq\ell$, and symmetric, so
we can replace $\veps{k_1k_2}(\Delta\ell_z)
\veps{h_1h_2}(\Delta\ell_z)^*$ there by
$\delta^{k_1h_1}\delta^{k_2h_2}$. The second line is just (\ref{eq:6j-2}).
\subsection{$\boldsymbol{s=3/2=s'}$}
\label{sec:3h}
The spin matrix elements entering (\ref{eq:CGmaster}) when
$s=3/2=s'$ are given by
\begin{equation}
\label{eq:3h-1}
\begin{aligned}
\veps{k}(\Delta\ell_z)^* \langle 3/2,s'_z | S^k | 3/2,s_z \rangle &=
\frac{3}{2} \veps{k}(\Delta\ell_z)^*
\vchiU{A}{i}(s'_z)^*\sigma^k_{AB}\vchiU{B}{i}(s_z)~,\\
\veps{ij}(\Delta\ell_z)^* \langle 3/2,s'_z | S^iS^j | 3/2,s_z \rangle &=
-3 \veps{ij}(\Delta\ell_z)^*
\vchiU{A}{i}(s'_z)^*\vchiU{A}{j}(s_z)~,\\
\veps{ijk}(\Delta\ell_z)^* \langle 3/2,s'_z | S^iS^jS^k | 3/2,s_z \rangle &=
-\frac{3}{2} \veps{ijk}(\Delta\ell_z)^*
\vchiU{A}{i}(s'_z)^* \sigma^j_{AB} \vchiU{B}{k}(s_z)~.
\end{aligned}
\end{equation}
The orbital angular momentum change is $0\le \Delta\ell \le 3$.
\paragraph*{$\boldsymbol{|\Delta\ell|=0}$}
\label{sec:3h-1}
Substituting (\ref{eq:3h-1}) in (\ref{eq:CGmaster}) we get
\begin{subequations}
\label{eq:3h-2}
\begin{equation}
\label{eq:3h-2a}
\begin{aligned}
S(j,\ell,\ell,3/2,3/2) &= C^{\frac{3}{2}\frac{3}{2}0}_{\ell\ell j} \delta_{\ell'_z
\ell_z} \delta_{s'_zs_z} + \frac{3}{2}C^{\frac{3}{2}\frac{3}{2}1}_{\ell\ell
j} \langle \ell,\ell'_z | \veps{i}(\Delta\ell_z )L^i | \ell,\ell_z
\rangle \veps{k}(\Delta\ell_z)^* \vchiU{A}{h}(s'_z)^*
\sigma^k_{AB} \vchiU{B}{h}(s_z)\\
&\quad -3 C^{\frac{3}{2}\frac{3}{2}2}_{\ell\ell j} \langle \ell,\ell'_z |
\veps{i_1i_2}(\Delta\ell_z)L^{i_1}L^{i_2} |\ell,\ell_z \rangle
\veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^* \vchiU{A}{h_2}(s_z)\\
&\quad -\frac{3}{2} C^{\frac{3}{2}\frac{3}{2}3}_{\ell\ell j} \langle \ell,\ell'_z |
\veps{i_1 i_2 i_3}(\Delta\ell_z) L^{i_1}L^{i_2}L^{i_3} | \ell,\ell_z
\rangle \veps{h_1h_2h_3}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchiU{B}{h_3}(s_z)~,
\end{aligned}
\end{equation}
where the coefficients $C^{\frac{3}{2}\frac{3}{2}\Delta}_{\ell\ell j}$
given in (\ref{eq:CGmaster}) in this case take the form
\begin{equation}
\label{eq:3h-2b}
\begin{aligned}
C^{\frac{3}{2}\frac{3}{2}0}_{\ell\ell j} &= \frac{1}{4}\,\frac{2j+1}{2\ell+1}~,\\
C^{\frac{3}{2}\frac{3}{2}1}_{\ell\ell j} &= \frac{2}{5}\, \frac{j-\ell}{2\ell+1}\,
\left(j+\frac{1}{2}+4 \left(j-\ell-\frac{3}{2}\right) (j-\ell)
\left(j-\ell+\frac{3}{2}\right)
\right)\frac{1}{\widetilde{D}_{\ell j}}~,\\
C^{\frac{3}{2}\frac{3}{2}2}_{\ell\ell j} &= -\frac{1}{2}\, \frac{1}{2\ell+1}\, \left(
\frac{5}{4} - (j-\ell)^2 \right) \left( j+\frac{1}{2}-2(j-\ell)^3 +
\frac{9}{2} (j-\ell) \right) \frac{1}{D_{\ell j}}~,\\
C^{\frac{3}{2}\frac{3}{2}3}_{\ell\ell j} &= -\frac{1}{2}\,
\frac{(-1)^{j-\ell-1/2}}{|j-\ell|(2\ell+1)}\, \frac{1}{D_{\ell j}}~,\\
\widetilde{D}_{\ell j} &= \left( \ell+1+\frac{2}{3}\left(j-\ell-\frac{3}{2}\right)
\left(j-\ell+\frac{1}{2}\right)(j-\ell+1) \right)~,\\
D_{\ell j} &= \widetilde{D}_{\ell j}\times
\left( j - \frac{2}{3}\left(j-\ell-\frac{3}{2}\right)
\left(j-\ell+\frac{1}{2}\right)(j-\ell+1) \right)~.
\end{aligned}
\end{equation}
\end{subequations}
As above, we notice that the matrix elements in (\ref{eq:3h-2a}) can
also be written as,
\begin{equation}
\label{eq:3halfaux1}
\begin{aligned}
\veps{p}(\Delta\ell_z) \langle\ell,\ell'_z|L^p|\ell,\ell_z\rangle
\veps{k}(\Delta\ell_z)^*
\vchiU{A}{i}(s'_z)^*\sigma^k_{AB}\vchiU{B}{i}(s_z)&=
\langle \ell,\ell'_z | L^k |
\ell,\ell_z \rangle
\vchiU{A}{i}(s'_z)^*\sigma^k_{AB}\vchiU{B}{i}(s_z)~, \\
2\veps{pq}(\Delta\ell_z)
\langle\ell,\ell'_z|L^pL^q|\ell,\ell_z\rangle
\veps{ij}(\Delta\ell_z)^* \vchiU{A}{i}(s'_z)^* \vchiU{B}{j}(s_z) &=
\langle\ell,\ell'_z|L^{\{i}L^{j\}_0}|\ell,\ell_z\rangle
\vchiU{A}{i}(s'_z)^* \vchiU{A}{j}(s_z) ~,\\
6\veps{pqr}(\Delta\ell_z)\langle\ell,\ell'_z|L^pL^qL^r|\ell,\ell_z\rangle
\veps{ijk}(\Delta\ell_z)^* \vchiU{A}{i}(s'_z)^*
\sigma^k_{AB}\vchiU{B}{j}(s_z)
&=
\langle \ell,\ell'_z | L^{\{i}L^{j}L^{k\}_0} |\ell,\ell_z \rangle
\\
&\qquad\times \vchiU{A}{i}(s'_z)^* \sigma^j_{AB}\vchiU{B}{k}(s_z) ~.
\end{aligned}
\end{equation}
With the matrix elements written as in (\ref{eq:3h-2a}),
$S(j,\ell,\ell,3/2,3/2)$ takes the form needed in the
derivation of addition theorems for spin-3/2 spherical harmonics.
Similarly, (\ref{eq:6j-2}) can be rewritten as
\begin{equation}
\label{eq:3h-3}
\begin{gathered}
\begin{aligned}
S(j,\ell,\ell,3/2,3/2) &= \kappa^{\frac{3}{2}\frac{3}{2}0}_{\ell\ell j} \delta_{\ell'_z \ell_z}
\delta_{s'_zs_z} + \kappa^{\frac{3}{2}\frac{3}{2}1}_{\ell\ell j} \CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{3/2}{s'_z}{1}{\Delta\ell_z}{3/2}{s_z}\\
&\quad +\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}\\
&\quad +\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell\ell j}\CG{\ell}{\ell_z}{3}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{3/2}{s'_z}{3}{\Delta\ell_z}{3/2}{s_z}~,
\end{aligned}\\
\kappa^{\frac{3}{2}\frac{3}{2}0}_{\ell\ell j} = C^{\frac{3}{2}\frac{3}{2}0}_{\ell\ell j},
\quad
\kappa^{\frac{3}{2}\frac{3}{2}1}_{\ell\ell j} = \sqrt{\frac{15}{4}} \sqrt{\ell(\ell+1)}
C^{\frac{3}{2}\frac{3}{2}1}_{\ell\ell j},
\quad
\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell\ell j} = \sqrt{\frac{5}{4}}
\sqrt{\ell(\ell+1)} \sqrt{(2\ell-1)(2\ell+3)}
C^{\frac{3}{2}\frac{3}{2}2}_{\ell\ell j},\\
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell\ell j} = \frac{3}{2} \sqrt{\frac{7}{20}}
\sqrt{(\ell-1)\ell(\ell+1)(\ell+2)} \sqrt{(2\ell-1)(2\ell+3)}
C^{\frac{3}{2}\frac{3}{2}3}_{\ell\ell j}~,
\end{gathered}
\end{equation}
Eqs.\ (\ref{eq:3h-2}) and (\ref{eq:3h-3}) are of course related by
the WE theorem, with the reduced matrix elements in appendix
\ref{sec:matele}.
\paragraph*{$\boldsymbol{|\Delta\ell|=1}$}
\label{sec:3h-2}
In this case $S(j,\ell',\ell,3/2,3/2)$ can be non-vanishing only if
$\ell'=\ell+1$ and $j=\ell+3/2$, $\ell+1/2$, $\ell-1/2$, or if
$\ell'=\ell-1$ and $j=\ell+1/2$, $\ell-1/2$, $\ell-3/2$.
Evaluating the spin matrix elements with (\ref{eq:3h-1}) we find,
\begin{subequations}
\label{eq:3h-4}
\begin{equation}
\label{eq:3h-4a}
\begin{aligned}
S(j,\ell_1,\ell,3/2,3/2) &=
\frac{3}{2}C^{\frac{3}{2}\frac{3}{2}1}_{\ell_1\ell j}
\langle\ell_1,\ell'_z | \veps{k}(\Delta\ell_z) \verr{k} |
\ell,\ell_z\rangle \, \veps{h}(\Delta\ell_z)^*
\vchiU{A}{j}(s'_z)^*\sigma^h_{AB}\vchiU{B}{j}(s_z) \\
&-3 C^{\frac{3}{2}\frac{3}{2}2}_{\ell_1\ell j}
\langle\ell_1,\ell'_z | \veps{i_1i_2}(\Delta\ell_z)\verr{i_1}L^{i_2} |
\ell,\ell_z\rangle \, \veps{j_1j_2}(\Delta\ell_z)^*
\vchiU{A}{j_1}(s'_z)^*
\vchiU{A}{j_2}(s_z)\\
&-\frac{3}{2}C^{\frac{3}{2}\frac{3}{2}3}_{\ell_1\ell j}
\langle\ell_1,\ell'_z | \veps{i_1i_2i_3}(\Delta\ell_z) \verr{i_1} L^{i_2}
L^{i_3}|\ell,\ell_z\rangle \, \veps{j_1j_2j_3}(\Delta\ell_z)^*
\vchiU{A}{j_1}(s'_z)^*\sigma^{j_2}_{AB}\vchiU{B}{j_3}(s_z)~,
\end{aligned}
\end{equation}
with the coefficients $C^{\frac{3}{2}\frac{3}{2}\Delta}_{\ell_1\ell j}$ given by
\begin{align}
C^{\frac{3}{2}\frac{3}{2}1}_{\ell_1\ell j} &= -\frac{1}{5}\,
\frac{j+1/2}{2\langle\ell\rangle+1}
\frac{\sqrt{\left(\rule{0pt}{9pt} j+\langle\ell\rangle+1\right)^2 - 4}\,
\sqrt{4 - \left(\rule{0pt}{9pt} j-\langle\ell\rangle\right)^2}}
{\sqrt{\langle\ell\rangle(\langle\ell\rangle+1)}}
~,\nonumber\\
C^{\frac{3}{2}\frac{3}{2}2}_{\ell_1\ell j} &= -2 \frac{\left(\rule{0pt}{10pt}
3\langle\ell\rangle-j+1\right)}
{(2\langle\ell\rangle+1)\sqrt{\langle\ell\rangle(\langle\ell\rangle+1)}}\,
\frac{ \left(\rule{0pt}{10pt}
(j-\langle\ell\rangle)(j+1/2)+1/2\right) \left(\rule{0pt}{10pt}
5(j-\langle\ell\rangle)^2-4\right)} {\sqrt{\left(
j+\langle\ell\rangle+1\right)^2 - 4} \sqrt{4 - \left(
j-\langle\ell\rangle\right)^2}}~, \label{eq:3h-4b}\\
C^{\frac{3}{2}\frac{3}{2}3}_{\ell_1\ell j} &=-4
\frac{\sqrt{\langle\ell\rangle(\langle\ell\rangle+1)}}{(2\langle\ell\rangle+1)j(j+1)}
\frac{\left(\rule{0pt}{10pt} 3 j-\langle\ell\rangle+1\right)
\left(\rule{0pt}{10pt} 3/2(j-\langle\ell\rangle)^2 -1\right)}
{\sqrt{\left( j+\langle\ell\rangle+1\right)^2 - 4} \sqrt{4 - \left(
j-\langle\ell\rangle\right)^2}}~.\nonumber
\end{align}
\end{subequations}
This expression in terms of matrix elements of irreducible tensor
operators is needed in the derivation of addition theorems for
spin-3/2 spherical harmonics. We notice also that the matrix elements
in (\ref{eq:3h-4a}) can be written without standard tensors as,
\begin{equation}
\label{eq:3h-4x}
\begin{gathered}
\langle\ell_1,\ell'_z | \veps{k}(\Delta\ell_z) \verr{k} |
\ell,\ell_z\rangle \, \veps{h}(\Delta\ell_z)^*
\vchiU{A}{j}(s'_z)^*\sigma^h_{AB}\vchiU{B}{j}(s_z) =
\langle\ell_1,\ell'_z | \verr{k} | \ell,\ell_z\rangle \,
\vchiU{A}{j}(s'_z)^*\sigma^k_{AB}\vchiU{B}{j}(s_z)~, \\
\langle\ell_1,\ell'_z | \veps{i_1i_2}(\Delta\ell_z)\verr{i_1}L^{i_2} |
\ell,\ell_z\rangle \, \veps{j_1j_2}(\Delta\ell_z)^*
\vchiU{A}{j_1}(s'_z)^* \vchiU{A}{j_2}(s_z) = \frac{1}{2}
\langle\ell_1,\ell'_z | \verr{\{i_1}L^{i_2\}} | \ell,\ell_z\rangle
\, \vchiU{A}{i_1}(s'_z)^* \vchiU{A}{i_2}(s_z)~,
\\
\begin{aligned}
\langle\ell_1,\ell'_z | \veps{i_1i_2i_3}(\Delta\ell_z) \verr{i_1} L^{i_2}
L^{i_3}|\ell,\ell_z\rangle \, \veps{j_1j_2j_3}(\Delta\ell_z)^*
\vchiU{A}{j_1}(s'_z)^*\sigma^{j_2}_{AB}\vchiU{B}{j_3}(s_z) &=
\frac{1}{6} \langle\ell_1,\ell'_z | \verr{\{i_1} L^{i_2}
L^{i_3\}_0}|\ell,\ell_z\rangle \\
&\quad\times
\vchiU{A}{i_1}(s'_z)^*\sigma^{i_2}_{AB}\vchiU{B}{i_3}(s_z)~.
\end{aligned}
\end{gathered}
\end{equation}
Analogously, (\ref{eq:6j-2}) can be
written as
\begin{subequations}
\label{eq:3h-5}
\begin{equation}
\label{eq:3h-5a}
\begin{aligned}
S(j,\ell_1,\ell,3/2,3/2) &= \kappa^{\frac{3}{2}\frac{3}{2}1}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{3/2}{s'_z}{1}{\Delta\ell_z}{3/2}{s_z} \\
&\quad+
\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}\\
&\quad+
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{3}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{3/2}{s'_z}{3}{\Delta\ell_z}{3/2}{s_z}~,
\end{aligned}
\end{equation}
with,
\begin{equation}
\label{eq:3h-5b}
\begin{aligned}
\kappa^{\frac{3}{2}\frac{3}{2}1}_{\ell_1\ell j} &= \frac{1}{2}
\sqrt{\frac{15}{2}} \frac{\Delta\ell}{\sqrt{2\ell_1+1}} \sqrt{2\langle\ell\rangle+1}
\, C^{\frac{3}{2}\frac{3}{2}1}_{\ell_1\ell j}~,\\
\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell_1\ell j} &= \frac{1}{4}
\sqrt{\frac{15}{2}} \frac{\Delta\ell}{\sqrt{2\ell_1+1}}
\sqrt{(2\langle\ell\rangle-1) (2\langle\ell\rangle+1)(2\langle\ell\rangle+3)}\,
C^{\frac{3}{2}\frac{3}{2}2}_{\ell_1\ell j}~, \\
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_1\ell j} &= \frac{1}{8}
\sqrt{\frac{21}{5}} \frac{\Delta\ell}{\sqrt{2\ell_1+1}}
\sqrt{(2\langle\ell\rangle-2)(2\langle\ell\rangle-1)(2\langle\ell\rangle+1)(2\langle\ell\rangle+3)(2\langle\ell\rangle+4)}\,
C^{\frac{3}{2}\frac{3}{2}3}_{\ell_1\ell j}~,
\end{aligned}
\end{equation}
\end{subequations}
where the coefficients $C^{\frac{3}{2}\frac{3}{2}\Delta}_{\ell_1\ell
j}$ are given in (\ref{eq:3h-4b}).
\paragraph*{$\boldsymbol{|\Delta\ell|=2}$}
\label{sec:3h-3}
In this case $S(j,\ell',\ell,3/2)$ can be non-vanishing only if
$\ell'=\ell+2$ and $j=\ell+3/2,\ell+1/2$ or $\ell'=\ell-2$ and
$j=\ell-1/2,\ell-3/2$. Taking into account this reduced set of $j$
values, and the spin matrix elements (\ref{eq:3h-1}), eq.\
(\ref{eq:CGmaster}) takes the form
\begin{equation}
\label{eq:3h-6}
\begin{gathered}
\begin{aligned}
S(j,\ell_2,\ell,3/2,3/2) &= -3
C^{\frac{3}{2}\frac{3}{2}2}_{\ell_2\ell j}
\langle\ell_2,\ell'_z|\veps{k_1
k_2}(\Delta\ell_z)\verr{k_1}\verr{k_2}|\ell, \ell_z\rangle
\veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\vchiU{A}{h_2}(s_z) \\
&- \frac{3}{2} C^{\frac{3}{2}\frac{3}{2}3}_{\ell_2\ell j}
\langle\ell_2,\ell'_z|\veps{k_1k_2k_3}(\Delta\ell_z)\verr{k_1}\verr{k_2}L^{k_3}|\ell,
\ell_z\rangle \veps{h_1h_2h_3}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchiU{B}{h_3}(s_z)~,
\end{aligned}\\
C^{\frac{3}{2}\frac{3}{2}2}_{\ell_2\ell j} = \frac{2}{\sqrt{3}}
\frac{3j-2\langle\ell\rangle+1/2}{\sqrt{(2j-1)(2j+3)}}~,
\quad
C^{\frac{3}{2}\frac{3}{2}3}_{\ell_2\ell j} = \frac{8}{\sqrt{3}}
\frac{j-\langle\ell\rangle}{\sqrt{(2j-1)(2j+3)}}~.
\end{gathered}
\end{equation}
We note here that the matrix elements appearing in this equation can
also be written as
\begin{equation}
\label{eq:3h-6aux}
\begin{gathered}
\langle\ell',\ell'_z|\veps{k_1
k_2}(\Delta\ell_z)\verr{k_1}\verr{k_2}|\ell, \ell_z\rangle
\veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\vchiU{A}{h_2}(s_z) =
\langle\ell',\ell'_z|\verr{i_1}\verr{i_2}|\ell, \ell_z\rangle
\vchiU{A}{i_1}(s'_z)^* \vchiU{A}{i_2}(s_z)~,\\
\begin{aligned}
\lefteqn{
\langle\ell',\ell'_z|\veps{k_1k_2k_3}(\Delta\ell_z)\verr{k_1}\verr{k_2}L^{k_3}|\ell,
\ell_z\rangle \veps{h_1h_2h_3}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchiU{B}{h_3}(s_z) =
}\hspace{48ex}\\
& \frac{1}{6}
\langle\ell',\ell'_z|\verr{\{i_1}\verr{i_2}L^{i_3\}_0}|\ell,
\ell_z\rangle \vchiU{A}{i_1}(s'_z)^*
\sigma^{i_2}_{AB}\vchiU{B}{i_3}(s_z)~,
\end{aligned}
\end{gathered}
\end{equation}
where the matrix element in the r.h.s.\ of the first equality is
traceless because $\ell'\neq \ell$. Similarly, from eq.\
(\ref{eq:6j-2}) we get
\begin{equation}
\label{eq:3h-7}
\begin{aligned}
S(j,\ell_2,\ell,3/2,3/2) &=
\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell_2\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_2}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}\\
&\quad+
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_2\ell j}
\CG{\ell}{\ell_z}{3}{\Delta\ell_z}{\ell_2}{\ell'_z}
\CG{3/2}{s'_z}{3}{\Delta\ell_z}{3/2}{s_z}~,\\
\kappa^{\frac{3}{2}\frac{3}{2}2}_{\ell_2\ell j} &=
\sqrt{\frac{15}{2}}\sqrt{2\ell+1} \sqrt{\frac{\langle\ell\rangle (\langle\ell\rangle+1)}
{(2\langle\ell\rangle-1)(2\langle\ell\rangle+1)(2\langle\ell\rangle+3)}}
C^{\frac{3}{2}\frac{3}{2}2}_{\ell_2\ell j} ~,\\
\quad
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_2\ell j} &= \frac{1}{2}
\sqrt{\frac{21}{2}} \sqrt{2\ell+1}
\sqrt{\frac{(\langle\ell\rangle-1)\langle\ell\rangle (\langle\ell\rangle+1)(\langle\ell\rangle+2)}
{(2\langle\ell\rangle-1)(2\langle\ell\rangle+1)(2\langle\ell\rangle+3)}}
C^{\frac{3}{2}\frac{3}{2}3}_{\ell_2\ell j} ~,
\end{aligned}
\end{equation}
with the coefficients $C^{\frac{3}{2}\frac{3}{2}\Delta}_{\ell_2\ell
j}$ of (\ref{eq:3h-6}).
\paragraph*{$\boldsymbol{|\Delta\ell|=3}$}
\label{sec:3h-4}
In this case $S(j,\ell',\ell,3/2,3/2)$ can be non-vanishing only if
$\ell'=\ell+3$ and $j=\ell+3/2$, or $\ell'=\ell-3$ and $j=\ell-3/2$.
Eq.\ (\ref{eq:CGmaster}) then reduces to
\begin{equation}
\label{eq:3h-8}
\begin{aligned}
S(j,\ell_3,\ell,3/2,3/2) &=
-\frac{3}{2}C^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j}
\langle\ell_3,\ell'_z|\veps{k_1k_2k_3}(\Delta\ell_z)\verr{k_1}\verr{k_2}\verr{k_3}|\ell,
\ell_z\rangle \\
&\quad \times\veps{h_1h_2h_3}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchiU{B}{h_3}(s_z)~, \\
C^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j} &= -\frac{8}{3}
\sqrt{\frac{j(j+1)}{(2j-1)(2j+3)}}~,
\end{aligned}
\end{equation}
which is the form needed to formulate an addition theorem for
spin-3/2 spherical harmonics with $\ell'=\ell\pm 3$. The matrix
element in (\ref{eq:3h-8}) can also be written
\begin{equation}
\label{eq:3h-9}
\begin{split}
\langle\ell',\ell'_z|\veps{k_1k_2k_3}(\Delta\ell_z)\verr{k_1}\verr{k_2}\verr{k_3}|\ell,
\ell_z\rangle
\veps{h_1h_2h_3}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchiU{B}{h_3}(s_z) =
\hspace{25ex}\\
=\frac{1}{6}
\langle\ell',\ell'_z|\verr{\{k_1}\verr{k_2}\verr{k_3\}_0}|\ell,
\ell_z\rangle
\vchiU{A}{k_1}(s'_z)^* \sigma^{k_2}_{AB}\vchiU{B}{k_3}(s_z)~.
\end{split}
\end{equation}
Similarly, (\ref{eq:6j-2}) reduces to
\begin{equation}
\label{eq:3h-a}
\begin{aligned}
S(j,\ell_3,\ell,3/2,3/2) &=
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j}
\CG{\ell}{\ell_z}{3}{\Delta\ell_z}{\ell_3}{\ell'_z}
\CG{3/2}{s'_z}{3}{\Delta\ell_z}{3/2}{s_z}~,\\
\kappa^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j} &=
\frac{\sqrt{7}}{32} \Delta\ell \sqrt{\frac{2\ell+1}{\langle\ell\rangle+2}}
\sqrt{\frac{(2\langle\ell\rangle-1) (2\langle\ell\rangle+1) (2\langle\ell\rangle+3)}{(\langle\ell\rangle-1) \langle\ell\rangle
(\langle\ell\rangle+1)}}C^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j} ~,
\end{aligned}
\end{equation}
with $C^{\frac{3}{2}\frac{3}{2}3}_{\ell_3\ell j}$ from (\ref{eq:3h-8}).
\subsection{The spin transition operator}
\label{sec:spintrans}
We define the spin transition operator $\vec{T}$ as a spin vector
operator, commuting with all orbital operators, and satisfying
\begin{equation}
\label{eq:Top}
[S^i,T^j] = i\varepsilon^{ijk} T^k~,
\qquad
\langle s' || \veps{i}T^i || s \rangle = \frac{s'-s}{\sqrt{2}}
\sqrt{\frac{2\langle s\rangle+1}{2s'+1}} \left(\delta_{s'(s+1)}+\delta_{s'(s-1)}\right).
\end{equation}
From (\ref{eq:Top}) we have,
\begin{equation}
\label{eq:Top2}
\langle s',s'_z | \vec{T}^2 | s, s_z\rangle =
\begin{cases}
\delta_{s's}\delta_{s'_zs_z}& \text{if $s'>1/2$ or $s>1/2$}\\
\frac{3}{4}\delta_{s'_zs_z}& \text{if $s'=s=1/2$}\
\end{cases}~.
\end{equation}
From (\ref{eq:Top}) and (\ref{eq:Top2}) the operator $\vec{T}$ is seen
to be the spin-space analog of the orbital operator $\verr{}$. In
particular, the reduced matrix elements of tensor products of
$\verr{}$ and $\vec{L}$ given in Appendix \ref{sec:matele} apply
without changes to tensor products of $\vec{T}$ and $\vec{S}$.
\subsection{$\boldsymbol{s',s=1,0}$ or $\boldsymbol{0,1}$}
\label{sec:10}
The relevant spin matrix element in this
case is,
\begin{equation}
\label{eq:10-1}
\langle 1, s'_z | T^h | 0, 0 \rangle = \frac{1}{\sqrt{3}}
\veps{h}(s'_z)^*~.
\end{equation}
$|\Delta\ell|$ can take the values 0 and 1.
\paragraph*{$\boldsymbol{|\Delta\ell|=0}$}
\label{sec:10-1}
$S(j,\ell,\ell,1,0)$ (resp.\ $S(j,\ell,\ell,0,1)$) can be non-vanishing only if
$j=\ell$ (resp.\ $j=\ell'$).
Then, (\ref{eq:CGmaster}) and (\ref{eq:6j-2}) take the form,
\begin{equation}
\label{eq:10-2}
\begin{gathered}
S(\ell,\ell,\ell,1,0) = \frac{1}{\sqrt{3}} C^{101}_{\ell\ell \ell} \langle
\ell, \ell'_z | L^k | \ell, \ell_z \rangle \veps{k}(s'_z)^*
= -\sqrt{3} \CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1}{s'_z}{1}{\Delta\ell_z}{0}{0}~,\\
C^{101}_{\ell\ell \ell} = \sqrt{\frac{3}{\ell(\ell+1)}}~.
\end{gathered}
\end{equation}
Similarly,
\begin{equation}
\label{eq:10-3}
S(\ell,\ell,\ell,0,1) = \frac{1}{\sqrt{3}} C^{011}_{\ell\ell \ell} \langle
\ell, \ell'_z | L^k | \ell, \ell_z \rangle \veps{k}(s_z)
= \CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{0}{0}{1}{\Delta\ell_z}{1}{s_z}~,
\end{equation}
with $C^{011}_{\ell\ell \ell} = C^{101}_{\ell\ell \ell}$.
\paragraph*{$\boldsymbol{|\Delta\ell|=1}$}
\label{sec:10-2}
In this case $S(j,\ell',\ell,s',s)$ can be non-vanishing only if
$j=\ell$ (if $s=0$) or $\ell'$ (if $s'=0$). Thus, with the spin
matrix element (\ref{eq:10-1}) we get, from (\ref{eq:CGmaster}) and
(\ref{eq:6j-2})~,
\begin{equation}
\label{eq:10-4}
\begin{gathered}
S(\ell,\ell_1,\ell,1,0) = \frac{C^{101}_{\ell_1\ell\ell}}{\sqrt{3}}
\langle \ell_1,\ell'_z | \verr{k} | \ell,\ell_z\rangle
\veps{k}(s'_z)^*
= \sqrt{3} \sqrt{\frac{2\ell+1}{2\ell_1+1}}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1}{s'_z}{1}{\Delta\ell_z}{0}{0}~,\\
C^{101}_{\ell_1\ell\ell} = -\sqrt{6} \Delta\ell
\sqrt{\frac{2\ell+1}{2\langle\ell\rangle+1}}~.
\end{gathered}
\end{equation}
Analogously,
\begin{equation}
\label{eq:10-5}
\begin{gathered}
S(\ell_1,\ell_1,\ell,0,1) = \frac{C^{011}_{\ell_1\ell\ell_1}}{\sqrt{3}}
\langle \ell_1,\ell'_z | \verr{k} | \ell,\ell_z\rangle
\veps{k}(s'_z)
=
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1}{s'_z}{1}{\Delta\ell_z}{0}{0}~,\\
C^{011}_{\ell_1\ell\ell'} = \sqrt{6} \Delta\ell
\sqrt{\frac{2\ell'+1}{2\langle\ell\rangle+1}}~.
\end{gathered}
\end{equation}
\subsection{$\boldsymbol{s',s=3/2,1/2}$ or
$\boldsymbol{1/2,3/2}$}
\label{sec:3h1h}
The spin matrix elements
appearing in (\ref{eq:CGmaster}) in this case are
\begin{equation}
\label{eq:3h1h-1}
\begin{aligned}
\langle 3/2,s'_z | \veps{h}(\Delta\ell_z)^* T^h | 1/2, s_z \rangle &=
\frac{1}{2} \sqrt{\frac{3}{2}} \veps{h}(\Delta\ell_z)^*
\vchiU{A}{h}(s'_z)^* \vchi{A}(s_z)~,\\
\langle 3/2,s'_z | \veps{h_1h_2}(\Delta\ell_z)^* T^{h_1}S^{h_2} | 1/2, s_z
\rangle &= \frac{1}{4} \sqrt{\frac{3}{2}} \veps{h_1h_2}(\Delta\ell_z)^*
\vchiU{A}{h_1}(s'_z)^* \sigma^{h_2}_{AB} \vchi{B}(s_z)~.
\end{aligned}
\end{equation}
There are three possible values for $0\le |\Delta\ell| \le 2$.
\paragraph*{$\boldsymbol{|\Delta\ell|=0}$}
\label{sec:3h1h-1}
If $s'=3/2$, $s=1/2$, and $\ell'=\ell$, then $S(j,\ell,\ell,3/2,1/2)$
can be non-vanishing only if $j=\ell\pm1/2$. In the case $s'=1/2$,
$s=3/2$, it must be $j=\ell'\pm1/2$.
Substituting (\ref{eq:3h1h-1}) in (\ref{eq:CGmaster}) we obtain
\begin{subequations}
\label{eq:3h1h-2}
\begin{equation}
\label{eq:3h1h-2a}
\begin{aligned}
S(j,\ell,\ell,3/2,1/2) &= \frac{1}{2} \sqrt{\frac{3}{2}}
C^{\frac{3}{2}\frac{1}{2}1}_{\ell\ell j} \langle \ell, \ell'_z | \veps{k}(\Delta\ell_z)
L^k | \ell, \ell_z \rangle \veps{h}(\Delta\ell_z)^* \vchiU{A}{h}(s'_z)^*
\vchi{A}(s_z) \\
&\quad +
\frac{1}{4} \sqrt{\frac{3}{2}} C^{\frac{3}{2}\frac{1}{2}2}_{\ell\ell
j} \langle \ell, \ell'_z | \veps{k_1 k_2}(\Delta\ell_z) L^{k_1} L^{k_2} |
\ell, \ell_z \rangle \veps{h_1h_2}(\Delta\ell_z)^*
\vchiU{A}{h_1}(s'_z)^* \sigma^{h_2}_{AB} \vchi{B}(s_z)~,
\end{aligned}
\end{equation}
with
\begin{equation}
\label{eq:3h1h-2b}
C^{\frac{3}{2}\frac{1}{2}1}_{\ell\ell j} = \frac{2}{2\ell+1}
\sqrt{\frac{2j-\ell+1/2}{2\ell-j+1/2}}~,
\qquad
C^{\frac{3}{2}\frac{1}{2}2}_{\ell\ell j} = 8 \frac{j-\ell}{2\ell+1}
\frac{1}{\sqrt{(2j-\ell+1/2)(2\ell-j+1/2)}}~.
\end{equation}
\end{subequations}
The matrix elements in (\ref{eq:3h1h-2a}) can also be written without
standard tensors as,
\begin{equation}
\label{eq:3h1h-3}
\begin{gathered}
\langle \ell, \ell'_z | \veps{k}(\Delta\ell_z) L^k | \ell, \ell_z \rangle
\veps{h}(\Delta\ell_z)^* \vchiU{A}{h}(s'_z)^* \vchi{A}(s_z) = \langle
\ell, \ell'_z | L^k | \ell, \ell_z \rangle \vchiU{A}{k}(s'_z)^*
\vchi{A}(s_z) ~,\\
\begin{aligned}
\lefteqn{
\langle \ell, \ell'_z | \veps{k_1 k_2}(\Delta\ell_z) L^{k_1} L^{k_2} |
\ell, \ell_z \rangle \veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB} \vchi{B}(s_z) = }\hspace{53ex}\\
& =\frac{1}{2}\langle \ell, \ell'_z |
L^{\{k_1} L^{k_2\}_0} | \ell, \ell_z \rangle
\vchiU{A}{k_1}(s'_z)^* \sigma^{k_2}_{AB}
\vchi{B}(s_z)~.
\end{aligned}
\end{gathered}
\end{equation}
Similarly, (\ref{eq:6j-2}) takes the simpler form,
\begin{equation}
\label{eq:3h1h-4}
\begin{split}
S(j,\ell,\ell,3/2,1/2) = -\frac{\sqrt{3}}{2} \sqrt{\ell(\ell+1)}
C^{\frac{3}{2}\frac{1}{2}1}_{\ell\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{3/2}{s'_z}{1}{\Delta\ell_z}{1/2}{s_z}
\hspace{20ex}\\
- \frac{\sqrt{5}}{8} \sqrt{\ell(\ell+1)} \sqrt{(2\ell-1)(2\ell+3)}
C^{\frac{3}{2}\frac{1}{2}2}_{\ell\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{1/2}{s_z}~,
\end{split}
\end{equation}
with the coefficients given in (\ref{eq:3h1h-2b}).
The case $s'=1/2$, $s=3/2$ is completely analogous. From
(\ref{eq:CGmaster}) we find
\begin{equation}
\label{eq:3h1h-5}
\begin{aligned}
S(j,\ell,\ell,1/2,3/2) &= \frac{1}{2} \sqrt{\frac{3}{2}}
C^{\frac{1}{2}\frac{3}{2}1}_{\ell\ell j} \langle \ell, \ell'_z | \veps{k}(\Delta\ell_z)
L^k | \ell, \ell_z \rangle \veps{h}(\Delta\ell_z)^* \vchi{A}(s'_z)^*
\vchiU{A}{h}(s_z) \\
&\quad +
\frac{1}{4} \sqrt{\frac{3}{2}} C^{\frac{1}{2}\frac{3}{2}2}_{\ell\ell
j} \langle \ell, \ell'_z | \veps{k_1 k_2}(\Delta\ell_z) L^{k_1} L^{k_2} |
\ell, \ell_z \rangle \veps{h_1h_2}(\Delta\ell_z)^*
\vchi{A}(s'_z)^* \sigma^{h_2}_{AB} \vchiU{B}{h_1}(s_z)~,
\end{aligned}
\end{equation}
with $C^{\frac{1}{2}\frac{3}{2}\Delta}_{\ell\ell j} =
C^{\frac{3}{2}\frac{1}{2}\Delta}_{\ell\ell j}$ (see
(\ref{eq:3h1h-2b})), and from
(\ref{eq:6j-2})
\begin{equation}
\label{eq:3h1h-6}
\begin{split}
S(j,\ell,\ell,1/2,3/2) = \frac{1}{2} \sqrt{\frac{3}{2}}
\sqrt{\ell(\ell+1)} C^{\frac{1}{2}\frac{3}{2}1}_{\ell\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1/2}{s'_z}{1}{\Delta\ell_z}{3/2}{s_z}
\hspace{15ex}\\
+ \frac{1}{8} \sqrt{\frac{5}{2}} \sqrt{\ell(\ell+1)}
\sqrt{(2\ell-1)(2\ell+3)}
C^{\frac{1}{2}\frac{3}{2}2}_{\ell\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell}{\ell'_z}
\CG{1/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}~.
\end{split}
\end{equation}
The expressions (\ref{eq:3h1h-2}) and (\ref{eq:3h1h-5}) will be used
below to derive addition theorems involving one spin-3/2 and one
spin-1/2 spherical harmonic.
\paragraph*{$\boldsymbol{|\Delta\ell|=1}$}
\label{sec:3h1h-2}
We consider the case $s'=3/2$, $s=1/2$ first. In this case
$S(j,\ell',\ell,3/2,1/2)$ can be non-vanishing only if
$\ell'=\ell+1$ and $j=\ell\pm1/2$, or if $\ell'=\ell-1$ and
$j=\ell\pm1/2$. With the spin matrix elements in (\ref{eq:3h1h-1}),
from (\ref{eq:CGmaster}) we obtain
\begin{subequations}
\label{eq:3h1h-7}
\begin{equation}
\label{eq:3h1h-7a}
\begin{aligned}
S(j,\ell_1,\ell,3/2,1/2) &= \frac{1}{2} \sqrt{\frac{3}{2}}
C^{\frac{3}{2}\frac{1}{2}1}_{\ell_1\ell j} \langle\ell_1\ell'_z |
\veps{k}(\Delta\ell_z)\verr{k} | \ell,\ell_z\rangle
\veps{h}(\Delta\ell_z)^* \vchiU{A}{h}(s'_z)^* \vchi{A}(s_z)\\
&+ \frac{1}{4} \sqrt{\frac{3}{2}}
C^{\frac{3}{2}\frac{1}{2}2}_{\ell_1\ell j} \langle\ell_1\ell'_z |
\veps{k_1k_2}(\Delta\ell_z)\verr{k_1} L^{k_2} | \ell,\ell_z\rangle
\veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchi{B}(s_z)~,
\end{aligned}
\end{equation}
with the coefficients
\begin{equation}
\label{eq:3h1h-7b}
\begin{aligned}
C^{\frac{3}{2}\frac{1}{2}1}_{\ell_1\ell j} &=-\sqrt{2} \Delta s \Delta\ell
\sqrt{\frac{2j+1}{2\langle\ell\rangle+1}} \sqrt{\frac{2j-\langle\ell\rangle+\Delta s\Delta\ell
+\frac{1}{2}} {2\langle\ell\rangle-\Delta s\Delta\ell +1}}
\sqrt{1-2\Delta s\Delta\ell (j-\langle\ell\rangle)}~,\\
C^{\frac{3}{2}\frac{1}{2}2}_{\ell_1\ell j} &=-\frac{8}{\sqrt{3}} \sqrt{
\frac{\frac{3}{2}+\Delta s\Delta\ell(j-\langle\ell\rangle)} {2\langle\ell\rangle-\Delta s\Delta\ell + 1}}\,
\frac{1+2\Delta s\Delta\ell (j-\langle\ell\rangle)} {\sqrt{j+\frac{1}{2}+\Delta s \Delta\ell}}~,
\end{aligned}
\end{equation}
\end{subequations}
where in this case $\Delta s=1$.
On the first line of (\ref{eq:3h1h-7a}) we can replace
$\veps{k}(\Delta\ell_z) \veps{h}(\Delta\ell_z)^*$ with $\delta^{kh}$, and on the
second line,
\begin{equation}
\label{eq:3h1h-8}
\begin{split}
\langle\ell'\ell'_z | \veps{k_1k_2}(\Delta\ell_z)\verr{k_1} L^{k_2} |
\ell,\ell_z\rangle \veps{h_1h_2}(\Delta\ell_z)^* \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchi{B}(s_z) = \hspace{35ex}\\
=\frac{1}{2}
\langle\ell'\ell'_z | \verr{\{k_1} L^{k_2\}_0} |
\ell,\ell_z\rangle \vchiU{A}{h_1}(s'_z)^*
\sigma^{h_2}_{AB}\vchi{B}(s_z)~.
\end{split}
\end{equation}
Similarly, from (\ref{eq:6j-2}) we get
\begin{equation}
\label{eq:3h1h-9}
\begin{split}
S(j,\ell_1,\ell,3/2,1/2) = -\frac{1}{2}\sqrt{\frac{3}{2}}\Delta\ell
\sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}} C^{\frac{3}{2}\frac{1}{2}1}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{3/2}{s'_z}{1}{\Delta\ell_z}{1/2}{s_z}
\hspace{12ex}\\
- \frac{\sqrt{30}}{32} \Delta\ell \sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}}
\sqrt{(2\langle\ell\rangle-1)(2\langle\ell\rangle+3)} C^{\frac{3}{2}\frac{1}{2}2}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{1/2}{s_z}~,
\end{split}
\end{equation}
with the coefficients (\ref{eq:3h1h-7b}).
In the case $s'=1/2$, $s=3/2$, $S(j,\ell',\ell,1/2,3/2)$ can be
non-vanishing only if $\ell'=\ell+1$ and $j=\ell+3/2$, $\ell+1/2$,
or if $\ell'=\ell-1$ and $j=\ell-1/2$, $\ell-3/2$. The treatment of
this case is completely analogous to the previous one. In terms of
matrix elements, from (\ref{eq:CGmaster}) we obtain
\begin{equation}
\label{eq:3h1h-a}
\begin{aligned}
S(j,\ell_1,\ell,1/2,3/2) &= \frac{1}{2} \sqrt{\frac{3}{2}}
C^{\frac{1}{2}\frac{3}{2}1}_{\ell_1\ell j}
\langle\ell_1\ell'_z | \veps{k}(\Delta\ell_z)\verr{k} | \ell,\ell_z\rangle
\veps{h}(\Delta\ell_z)^* \vchi{A}(s'_z)^* \vchiU{A}{h}(s_z)\\
&+ \frac{1}{4} \sqrt{\frac{3}{2}}
C^{\frac{1}{2}\frac{3}{2}2}_{\ell_1\ell j}
\langle\ell_1\ell'_z | \veps{k_1k_2}(\Delta\ell_z) \verr{k_1} L^{k_2} |
\ell,\ell_z\rangle \veps{h_1h_2}(\Delta\ell_z)^* \vchi{A}(s'_z)^*
\sigma^{h_1}_{AB}\vchiU{B}{h_2}(s_z)~,
\end{aligned}
\end{equation}
with the coefficients $C^{\frac{1}{2}\frac{3}{2}\Delta}_{\ell_1\ell
j}$ given by (\ref{eq:3h1h-7b}) with $\Delta s=-1$. From (\ref{eq:6j-2})
we get, analogously,
\begin{equation}
\label{eq:3h1h-b}
\begin{split}
S(j,\ell_1,\ell,1/2,3/2) = \frac{\sqrt{3}}{4}\Delta\ell
\sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}} C^{\frac{1}{2}\frac{3}{2}1}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{1}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1/2}{s'_z}{1}{\Delta\ell_z}{3/2}{s_z}
\hspace{12ex}\\
+ \frac{\sqrt{15}}{32} \Delta\ell \sqrt{\frac{2\langle\ell\rangle+1}{2\ell_1+1}}
\sqrt{(2\langle\ell\rangle-1)(2\langle\ell\rangle+3)} C^{\frac{1}{2}\frac{3}{2}2}_{\ell_1\ell j}
\CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_1}{\ell'_z}
\CG{1/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}~.
\end{split}
\end{equation}
The equalities (\ref{eq:3h1h-7}) and (\ref{eq:3h1h-a}) will
be used for the derivation of an addition theorem involving one
spin-3/2 and one spin-1/2 spherical harmonic, with orbital angular
momentum differing by one unit.
\paragraph*{$\boldsymbol{|\Delta\ell|=2}$}
\label{sec:3h1h-3}
As in the previous paragraph, we consider the case $s'=3/2$, $s=1/2$
first. In this case $S(j,\ell',\ell,3/2,1/2)$ can be non-vanishing
only if $\ell'=\ell+2$ and $j=\ell+1/2$, or if $\ell'=\ell-2$ and
$j=\ell-1/2$. Using the spin matrix elements (\ref{eq:3h1h-1}) and
the range of $j$, (\ref{eq:CGmaster}) reduces to
\begin{subequations}
\label{eq:3h1h-c}
\begin{align}
\label{eq:3h1h-ca}
S(j,\ell_2,\ell,3/2,1/2) &= \frac{1}{4} \sqrt{\frac{3}{2}}
C^{\frac{3}{2}\frac{1}{2}2}_{\ell_2\ell j}
\langle\ell_2\ell'_z | \veps{k_1k_2}(\Delta\ell_z) \verr{k_1} \verr{k_2} |
\ell,\ell_z\rangle \veps{h_1h_2}(\Delta\ell_z)^*
\vchiU{A}{h_1}(s'_z)^* \sigma^{h_2}_{AB} \vchi{B}(s_z)~, \\
\label{eq:3h1h-cb}
C^{\frac{3}{2}\frac{1}{2}2}_{\ell_2\ell j} &= \frac{2}{\sqrt{3}} \Delta s
\Delta\ell \sqrt{2j+1} \sqrt{\frac{2\langle\ell\rangle+1}{\langle\ell\rangle(\langle\ell\rangle+1)}}~,
\end{align}
\end{subequations}
where $\Delta s=1$. The matrix elements in (\ref{eq:3h1h-ca}) can be
written without standard tensors as,
\begin{equation}
\label{eq:3h1h-cx}\hspace*{-7pt}
\langle\ell_2\ell'_z | \veps{k_1k_2}(\Delta\ell_z) \verr{k_1} \verr{k_2} |
\ell,\ell_z\rangle \veps{h_1h_2}(\Delta\ell_z)^*
\vchiU{A}{h_1}(s'_z)^* \sigma^{h_2}_{AB} \vchi{B}(s_z) =
\langle\ell_2\ell'_z | \verr{k_1} \verr{k_2} |
\ell,\ell_z\rangle
\vchiU{A}{k_1}(s'_z)^* \sigma^{k_2}_{AB} \vchi{B}(s_z).
\end{equation}
Similarly, in this case (\ref{eq:6j-2}) takes the
simplified form
\begin{equation}
\label{eq:3h1h-d}
\begin{aligned}
S(j,\ell_2,\ell,3/2,1/2) &= -\frac{\sqrt{30}}{8} \sqrt{2\ell+1}
\sqrt{\frac{\langle\ell\rangle (\langle\ell\rangle+1)}{(2\langle\ell\rangle-1)(2\langle\ell\rangle+1)(2\langle\ell\rangle+3)}}
C^{\frac{3}{2}\frac{1}{2}2}_{\ell_2\ell j} \\
&\quad\times \CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_2}{\ell'_z}
\CG{3/2}{s'_z}{2}{\Delta\ell_z}{1/2}{s_z}~.
\end{aligned}
\end{equation}
In the case $s'=1/2$, $s=3/2$, $S(j,\ell',\ell,1/2,3/2)$ can be
non-vanishing only if $\ell'=\ell+2$ and $j=\ell'-1/2$, or if
$\ell'=\ell-2$ and $j=\ell'+1/2$. In this case (\ref{eq:CGmaster})
reduces to
\begin{equation}
\begin{aligned}
\label{eq:3h1h-e}
S(j,\ell_2,\ell,1/2,3/2) &= \frac{1}{4} \sqrt{\frac{3}{2}}
C^{\frac{1}{2}\frac{3}{2}2}_{\ell_2\ell j}
\langle\ell_2\ell'_z | \veps{k_1k_2}(\Delta\ell_z) \verr{k_1} \verr{k_2} |
\ell,\ell_z\rangle \veps{h_1h_2}(\Delta\ell_z)^*
\vchi{A}(s'_z)^* \sigma^{h_2}_{AB} \vchiU{B}{h_1}(s_z)~, \\
\end{aligned}
\end{equation}
with $C^{\frac{1}{2}\frac{3}{2}2}_{\ell_2\ell j}$ given by the same
expression (\ref{eq:3h1h-cb}) as
$C^{\frac{3}{2}\frac{1}{2}2}_{\ell_2\ell j}$, with $\Delta s=-1$. The
equality (\ref{eq:6j-2}) takes the simplified form
\begin{equation}
\begin{aligned}
\label{eq:3h1h-f}
S(j,\ell_2,\ell,1/2,3/2) &= \frac{\sqrt{15}}{8} \sqrt{2\ell+1}
\sqrt{\frac{\langle\ell\rangle(\langle\ell\rangle+1)}{(2\langle\ell\rangle-1)
(2\langle\ell\rangle+1)(2\langle\ell\rangle+3)}}C^{\frac{1}{2}\frac{3}{2}2}_{\ell_2\ell j} \\
&\times \CG{\ell}{\ell_z}{2}{\Delta\ell_z}{\ell_2}{\ell'_z}
\CG{1/2}{s'_z}{2}{\Delta\ell_z}{3/2}{s_z}~,
\end{aligned}
\end{equation}
\section{Angular momentum projection operators}
\label{sec:proj}
We define the angular momentum projection operators as
\begin{equation}
\label{eq:proj}
\P_\ell = \sum_{m=-\ell}^\ell |\ell,m\rangle\langle\ell,m|~,
\qquad
0\leq\ell<\infty~.
\end{equation}
It is immediate that $\P_\ell$ are a sequence of orthogonal
projectors, which commute with $\vec{L}$ and resolve the identity,
\begin{equation}
\label{eq:id}
\P_{\ell'}\P_\ell = \P_\ell\delta_{\ell'\ell}~,
\qquad
\sum_{\ell=0}^\infty \P_\ell = 1~,
\qquad
[\vec{L},\P_\ell]=0~.
\end{equation}
The matrix elements of $\P_\ell$ in configuration space follow from
their definition (\ref{eq:proj}) and the addition theorem for
spherical harmonics
\begin{equation}
\label{eq:proj3}
\P_\ell|\verr{}\rangle = \sum_{m=-\ell}^\ell Y_{\ell
m}(\verr{})^*|\ell,m\rangle~,
\qquad
\langle\verrp{}|\P_\ell|\verr{}\rangle = \frac{2\ell+1}{4\pi}
P_\ell(\verr{}\cdot\verrp{})~.
\end{equation}
From (\ref{eq:proj})--(\ref{eq:proj3}) the matrix elements of
$\P_\ell$ with other orbital operators can be computed. Those matrix
elements, together with the results on products of CG coefficients of
the previous section, are the basic building blocks needed in our
derivation of addition theorems for spin spherical harmonics. As
will be discussed in II, however, the results for $\P_\ell$
matrix elements constitute by themselves addition theorems for
spherical harmonics.
\subsection{Matrix elements with tensor powers of $\boldsymbol{\vec{L}}$}
\label{sec:Lpower}
The matrix elements of $\P_\ell$ with tensor powers of $\vec{L}$ will
be used in the derivation of addition theorems for spin spherical
harmonics. For the irreducible components, from which the full tensor
matrix elements can be reconstructed, we have
\begin{equation}
\label{eq:recL}
\begin{aligned}
\langle\verrp{}|L^{\{h_1}\ldots L^{h_p\}_0} \P_\ell|\verr{}\rangle &=
\frac{2\ell+1}{4\pi}
i^p \left(\verr{}\wedge\nabla\right)^{\{h_p}
\ldots \left(\verr{}\wedge\nabla \right)^{h_1\}_0}
P_\ell(x) \\
&= i \left(\verr{}\wedge\nabla\right)^{\{h_p}
\langle\verrp{}|L^{h_1}\ldots L^{h_{p-1}\}_0}
\P_\ell|\verr{}\rangle \\
&=\frac{i}{(p-1)!} \left(\verr{}\wedge\nabla\right)^{\{h_p}
\langle\verrp{}|L^{\{h_1}\ldots L^{h_{p-1}\}_0\}_0}
\P_\ell|\verr{}\rangle~.
\end{aligned}
\end{equation}
Thus, we can obtain these irreducible matrix elements recursively,
with the result
\begin{subequations}
\label{eq:genL}
\begin{equation}
\label{eq:genLa}
\langle\verrp{}|L^{\{h_1}\ldots L^{h_p\}_0} \P_\ell|\verr{}\rangle
= \frac{2\ell+1}{4\pi} i^p \sum_{q=0}^{[p/2]}
C_{p,q} Z^{\{h_1\ldots h_q;h_{q+1}\ldots h_{2q}}
v^{h_{2q+1}}\ldots v^{h_p\}_0} P_\ell^{(p-q)}(x),
\end{equation}
with $P_\ell^{(k)}$ the $k^\mathrm{th}$ derivative of $P_\ell$ and,
\begin{equation}
\label{eq:genLb}
\begin{gathered}
x=\verr{}\cdot\verrp{}~,\quad \vec{v} = \verr{}\wedge\verrp{}~,\quad
C_{p,q} = ( 2q-1)!! \binom{p}{2q}~,\quad
C_{p,0} = 1~,\\
Z^{i_1\ldots i_{n_1};j_1\ldots j_{n_2}} = \verr{i_1} \ldots
\verr{i_{n_1}} \verrp{j_1} \ldots
\verrp{j_{n_2}} ~.
\end{gathered}
\end{equation}
For notational simplicity in what follows we adopt the following
useful conventions. In $Z^{i_1\ldots i_{n_1};j_1\ldots j_{n_2}}$ one
or both sets of indices may be empty,
\begin{equation}
\label{eq:genLbaux}
Z^{\mbox{ };j_1\ldots j_{n_2}} = \verrp{j_1} \ldots
\verrp{j_{n_2}}~,
\quad
Z^{i_1\ldots i_{n_2};\mbox{ }} = \verr{i_1} \ldots
\verr{i_{n_1}}~,
\quad
Z^{\mbox{ };\mbox{ }} = 1~.
\end{equation}
In the term $q=0$ in (\ref{eq:genLa}) the index sets are empty,
$Z^{h_1\ldots h_0;h_{1}\ldots h_{0}} \equiv Z^{\mbox{ };\mbox{ }}$.
Similarly, in the term $q=[p/2]$, when $p$ is even, we set
$v^{h_{2q+1}}\ldots v^{h_p} = v^{h_{p+1}}\ldots v^{h_p} \equiv 1$.
\end{subequations}
The cases $n=1$, 2, 3 will be needed for the derivation of addition
theorems for spherical harmonics of spin 1/2, 1 and 3/2 so we quote
them here explicitly,
\begin{subequations}
\label{eq:Ls}
\begin{align}
\label{eq:L}
\langle\verrp{}|\vec{L}\P_\ell|\verr{}\rangle &= i
\frac{2\ell+1}{4\pi} \vec{v} P'_\ell(\verr{}\cdot\verrp{})~, \\
\label{eq:LL}
\langle\verrp{}|L^{\{i} L^{j\}_0} \P_\ell|\verr{}\rangle
&= -\frac{2\ell+1}{4\pi} \left( v^{\{i} v^{j\}_0} P''_\ell(\verr{}\cdot\verrp{})
+ \verr{\{i} \verrp{j\}_0} P'_\ell(\verr{}\cdot\verrp{})
\right)~,\\
\label{eq:LLL}
\langle\verrp{}|L^{\{i} L^{j} L^{k\}_0} \P_\ell|\verr{}\rangle
&= -i \frac{2\ell+1}{4\pi} \left( v^{\{i} v^{j} v^{k\}_0}
P'''_\ell(\verr{}\cdot\verrp{}) + 3 \verrp{\{i} \verr{j} v^{k\}_0}
P''_\ell(\verr{}\cdot\verrp{}) \right)~.
\end{align}
\end{subequations}
\subsection{Matrix elements with tensor powers of $\boldsymbol{\verr{}}$}
\label{sec:rpower}
The matrix element $\langle
\verpp{}|\P_{\ell'}\verr{i_1}\ldots\verr{i_n}\P_\ell | \verp{}\rangle$
is an irreducible tensor if $\ell'=\ell\pm n$, otherwise reducible.
The irreducible matrix elements will be used in the derivation of
addition theorems for spin spherical harmonics. They satisfy the
recursion relation,
\begin{multline}
\label{eq:rmat1}
\langle\verpp{} |
\P_{\ell_{n+1}}\verr{i_1}\ldots\verr{i_{n+1}}\P_\ell |
\verp{}\rangle =\\
-\frac{1}{2\ell_n+1} \left[ (\ell_{n+1}-\ell_n)
\nabla^{i_{n+1}}_{p'} - \frac{1}{2} (\ell_{n+1} + \ell_{n}
+1) \verpp{k} \right]
\langle\verpp{}|\P_{\ell_{n}}\verr{i_1}\ldots\verr{i_{n}}\P_\ell |
\verp{}\rangle~,
\end{multline}
with
\begin{equation}
\label{eq:rmat2}
\langle\verpp{} |
\P_{\ell_{1}}\verr{}\P_\ell |\verp{}\rangle =
\frac{\ell_1 - \ell}{4\pi} \left( \verpp{} P'_{\ell_1}(x) - \verp{}
P'_\ell(x) \right)~.
\end{equation}
In these equations we used the notation $\ell_k = \ell\pm k$, see
appendix \ref{sec:notatio}. Eqs.\ (\ref{eq:rmat1}), (\ref{eq:rmat2})
are established in appendix \ref{sec:appa}.
The recursion relation (\ref{eq:rmat1}) can be solved with the initial
condition (\ref{eq:rmat2}) to give
\begin{equation}
\label{eq:rmat3}
\begin{gathered}
\langle\verpp{}|\P_{\ell_{n}}\verr{i_1}\ldots\verr{i_{n}}\P_\ell |
\verp{}\rangle = A_n
\sum^n_{\substack{k_1,k_2=0\\k_1+k_2=n}} \frac{(-1)^{k_2}}{k_1!k_2!}
Z^{\{i_1\ldots i_{k_2};i_{k_2+1}\ldots i_n\}_0} P_{\ell_{k_1}}^{(n)}(x)~,\\
A_n = \frac{(\ell_1-\ell)^n}{4\pi} \frac{2\ell+1}{\prod_{k=0}^{n-1}
(2\ell_k+1)}
= \frac{1}{4\pi} \left(\frac{\ell_n-\ell}{n}\right)^n (2\ell+1)
\frac{(\ell_n+\ell-n-\frac{\ell_n-\ell}{n})!!}
{(\ell_n+\ell+n-\frac{\ell_n-\ell}{n})!!}~,
\end{gathered}
\end{equation}
with $Z^{i_1\ldots i_{n_1};j_{1}\ldots j_{n_2}}$ defined in
(\ref{eq:genLb}), (\ref{eq:genLbaux}). As before, we adopt the
convention that in the term with $k_1=n$, $k_2=0$ in (\ref{eq:rmat3})
the first index set in $Z$ is empty, $Z^{i_1\ldots i_0; i_1\ldots i_n}
\equiv Z^{\mbox{ }; i_1\ldots i_n}$ and, in the term with $k_1=0$,
$k_2=n$, the second index set is empty $Z^{i_1\ldots i_n;
i_{n+1}\ldots i_n} \equiv Z^{i_1\ldots i_n; \mbox{ }}$.
The particular cases $n=1$, 2, 3 will be needed for the derivation of
addition theorems for spherical harmonics of spin 1/2, 1 and 3/2, so
we give them here explicitly,
\begin{subequations}
\label{eq:rmat4}
\begin{align}
\label{eq:rmat4a}
\langle\verpp{}|\P_{\ell_{1}}\verr{i}\P_\ell |\verp{}\rangle &=
\frac{\ell_1-\ell}{4\pi} \left(-\verp{i} P'_\ell(x) + \verpp{i}
P'_{\ell_1}(x) \right),\\
\label{eq:rmat4b}
\langle\verpp{}|\P_{\ell_{2}}\verr{i_1}\verr{i_{2}}\P_\ell |
\verp{}\rangle &=
\frac{1}{4\pi} \frac{1}{2\ell_1+1} \left(
\frac{1}{2} \verp{\{i_1} \verp{i_2\}_0} P''_\ell(x) -
\verp{\{i_1} \verpp{i_2\}_0} P''_{\ell_1}(x)
+\frac{1}{2} \verpp{\{i_1} \verpp{i_2\}_0} P''_{\ell_2}(x)
\right),\\
\nonumber
\langle\verpp{}|\P_{\ell_{3}}\verr{i_1}\verr{i_{2}}
\verr{i_{3}}\P_\ell | \verp{}\rangle &=
\frac{\ell_1-\ell}{4\pi} \frac{1}{(2\ell_2+1)(2\ell_1+1)} \left(
-\frac{1}{6} \verp{\{i_1} \verp{i_2} \verp{i_3\}_0} P'''_\ell(x)
+\frac{1}{2} \verp{\{i_1} \verp{i_2} \verpp{i_3\}_0}
P'''_{\ell_1}(x)\right.\\
\label{eq:rmat4c}
&\quad \left.
-\frac{1}{2} \verp{\{i_1} \verpp{i_2} \verpp{i_3\}_0}
P'''_{\ell_2}(x)
+\frac{1}{6} \verpp{\{i_1} \verpp{i_2} \verpp{i_3\}_0} P'''_{\ell_3}(x)
\right).
\end{align}
\end{subequations}
\subsection{Mixed matrix elements}
\label{sec:mix}
We consider now matrix elements of $\P_\ell$ with tensor products of
both $\vec{L}$ and $\verr{}$. The matrix elements of the irreducible
tensor operator $\verr{\{i_1}\ldots\verr{i_n} L^{h_1}\ldots
L^{h_p\}_0}$ are of the form
\begin{equation}
\label{eq:mix1}
\langle\verpp{} | \P_{\ell_{n}}\verr{\{i_1}\ldots\verr{i_{n}}
L^{h_1}\ldots L^{h_t\}_0}\P_\ell |\verp{}\rangle = i^t
(\verp{}\wedge\nabla_p)^{\{h_t} \ldots (\verp{}\wedge\nabla_p)^{h_1}
\langle\verpp{} | \P_{\ell_{n}}\verr{i_1}\ldots\verr{i_{n}\}_0}
\P_\ell |\verp{}\rangle ~.
\end{equation}
From this equation and (\ref{eq:rmat3}), for $t=1$ we obtain
\begin{equation}
\label{eq:mix2}
\langle\verpp{}|\P_{\ell_{n}}\verr{\{i_1}\ldots\verr{i_{n}}L^{h\}_0}
\P_\ell | \verp{}\rangle = i A_n
\sum^n_{\substack{k_1,k_2=0\\k_1+k_2=n}} (-1)^{k_2} \binom{n}{k_1}
Z^{\{i_1\ldots i_{k_2};i_{k_2+1}\ldots i_n}
v^{h\}_0}P_{\ell_{k_1}}^{(n+1)}(x)~.
\end{equation}
For $s>1$, equation (\ref{eq:mix1}) leads to the recursion relation
\begin{equation}
\label{eq:mix3}
\begin{split}
\langle\verpp{} | \P_{\ell_{n}}\verr{\{i_1}\ldots\verr{i_{n}}
L^{h_1}\ldots L^{h_t\}_0}\P_\ell |\verp{}\rangle = \hspace{60ex}\\
=\frac{i}{(n+t-1)!}
(\verp{}\wedge\nabla_p)^{\{h_t}
\langle\verpp{} | \P_{\ell_{n}}\verr{\{i_1}\ldots\verr{i_{n}}
L^{h_1}\ldots L^{h_{t-1}\}_0\}_0} \P_\ell |\verp{}\rangle ~,
\end{split}
\end{equation}
which can be solved with the initial condition (\ref{eq:mix2}) to
yield
\begin{equation}
\label{eq:mix4}
\begin{split}
\langle\verpp{} | \P_{\ell_{n}}\verr{\{i_1}\ldots\verr{i_{n}}
L^{h_1}\ldots L^{h_t\}_0}\P_\ell |\verp{}\rangle = \hspace{63ex}\\
=i^t A_n
\sum^n_{\substack{k_1,k_2=0\\k_1+k_2=n}} (-1)^{k_2} \binom{n}{k_1}
\sum_{q=0}^{[t/2]} C_{t,q}
Z^{\{i_1\ldots i_{k_2}h_1\ldots h_q;i_{k_2+1}\ldots i_n
h_{q+1}\ldots h_{2q}} v^{h_{2q+1}}\ldots v^{h_t\}_0}
P_{\ell_{k_1}}^{(n+t-q)}(x)~,
\end{split}
\end{equation}
with $[\mbox{ }]$ in the inner summation denoting integer part, $A_n$
as defined in (\ref{eq:rmat3}), and $C_{s,q}$, $x$, $\vec{v}$, and
$Z^{i_1\ldots i_{n_1};j_{1}\ldots j_{n_2}}$ as defined in
(\ref{eq:genLb}), with $\verp{}$ instead of $\verr{}$. As before, for
notational simplicity, we have not separated from the sum the terms
with $(k_1,k_2)=(n,0)$, $(0,n)$, those with $q=0$ and, for even $s$,
those with $q=[s/2]$. In those cases we apply the same conventions as
explained after eqs.\ (\ref{eq:genL}) and (\ref{eq:rmat3}).
Particular cases of importance for the derivation of addition theorems
for spherical harmonics of spin 1/2, 1 and 3/2, are
\begin{subequations}
\label{eq:mix5}
\begin{align}
\langle\verpp{} | \P_{\ell_{1}}\verr{\{i}L^{k\}}\P_\ell |\verp{}\rangle
&= i \frac{\ell_1-\ell}{4\pi} \left(\verpp{\{i}v^{k\}}
P_{\ell_1}''(x) - \verp{\{i}v^{k\}} P_{\ell}''(x)
\right),\label{eq:mix5a}\\
\langle\verpp{}|\P_{\ell_{2}}\verr{\{i}\verr{j}L^{k\}_0}
\P_\ell | \verp{}\rangle &= \frac{i}{4\pi} \frac{1}{2\ell_1+1}
\left(
\verpp{\{i}\verpp{j}v^{k\}_0} P'''_{\ell_2}(x) -
2\verpp{\{i}\verp{j}v^{k\}_0} P'''_{\ell_1}(x) +
\verp{\{i}\verp{j}v^{k\}_0} P'''_{\ell}(x)
\right),\label{eq:mix5b}\\
\langle\verpp{} | \P_{\ell_{1}}\verr{\{i}L^{j_1}L^{j_2\}_0}\P_\ell |\verp{}\rangle
&= - \frac{\ell_1-\ell}{4\pi} \left(\verpp{\{i}v^{j_1}v^{j_2\}_0}
P_{\ell_1}'''(x) - \verp{\{i}v^{j_1}v^{j_2\}_0} P_{\ell}'''(x)
+ \verpp{\{i}\verpp{j_1}\verp{j_2\}_0} P''_{\ell_1}(x)
\right. \nonumber\\
&\qquad\left. -
\verpp{\{i}\verp{j_1}\verp{j_2\}_0} P''_{\ell}(x) \right).\label{eq:mix5c}
\end{align}
\end{subequations}
\section{Final remarks}
\label{sec:finrem}
We have presented in the foregoing sections preliminary results needed
for the systematic derivation of addition theorems for spin spherical
harmonics in II. In sect.\ \ref{sec:fac} we obtained the
factorization of orbital and spin degrees of freedom in products of CG
coefficients of the form (\ref{eq:additionintro}) in general form, and
discussed the particular cases with $0\leq s',s\leq 3/2$. In those
cases the coefficients $C^{s's\Delta}_{\ell'\ell j}$, given in full
generality in (\ref{eq:CGmaster}), were reduced to much smaller forms,
and the tensor and spinor structures of each term in the expansion
given explicitly.
In sect.\ \ref{sec:proj} the matrix elements of the angular-momentum
projector operator with the irreducible components of arbitrary tensor
products of $\verr{}$ and $\vec{L}$ are given in general form, for
all values of their parameters. Those matrix elements will be used in
II to obtain general expressions for bilocal spherical harmonics, and
to derive addition theorems for spin spherical harmonics. Their
applicability is even wider, however, when they are appropriately
combined to obtain matrix elements of reducible tensor operators.
Some examples of those applications will also be considered in II.
As a side remark we point out that an unexpected byproduct of the
results of sect.\ \ref{sec:fac} is an improvement in computational
efficiency, at least in the specific case of infinite-precision
computation \cite{wolf} in which the results are given as a rational
number times the square root of a rational number. Consider, for
example, the computation of both sides of (\ref{eq:3h-3}), at fixed
$\ell$ and $j$, for all possible values of $-\ell\leq
\ell'_z,\ell_z\leq \ell$ and $-3/2\leq s'_z, s_z\leq 3/2$. If the
l.h.s.\ of (\ref{eq:3h-3}) is computed as written on the second line
of (\ref{eq:additionintro}), we find that the ratio of CPU times
$\tau_\mathrm{r.h.s.}/\tau_\mathrm{l.h.s.}$ begins at $\sim 2$ at
$\ell=1$, monotonically decreasing with $\ell$ to reach 1 at $\ell\sim
20$, $\sim 0.5$ at $\ell\sim100$, and $\sim 0.25$ at $\ell\sim200$.
We remark that those numbers are subject to statistical fluctuations.
|
\section{Introduction}
~~~The current state-of-the-art understanding on the subject of 4D, $\cal N $ $=$ 1 integer higher
spin supersymmetric multiplets was established in a work by Kuzenko and Sibiryakov \cite{Off2}
(KS) wherein they gave two such formulations for each and every possible value of the integer
superspin $Y$. These formulations are based on the introduction of constrained compensating
superfields. This seminal work laid a foundation for a number of latter studies \cite{Related}.
The goal of this work is to re-examine these schemes in order to be able to reproduce
their results and, if possible, to discover new formulations in the case of integer superspins. This is
exactly what will happen in the following. Their results will emerge naturally from our algorithm as
a possible way a theory of higher, integer massless superspins can be formulated. Also we will
discover the KS description is not the only consistent formulation and at least one alternative exists.
We approach the problem from a different angle, by studying actions for spinorial superfields.
This is supported from the naive observation that for a massless, integer higher superspin theory the
higher spin projection operator acting on the main superfield of the theory must give rise to a chiral
object with an even number of indices.
\begin{equation}
\left(\Pi\Psi\right)_{{\alpha}(2s-1)}\propto {\rm D}^{{\alpha}_{2s}}\ {\bar{\rm D}}^2{\rm D}_{({\alpha}_{2s}}\pa^{{\dot{\alpha}}_1}{}_{{\alpha}_{2s-1}}
\dots\pa^{{\dot{\alpha}}_{s-1}}{}_{{\alpha}_{s+1}}\Psi_{{\alpha}(s)){\dot{\alpha}}(s-1)} ~~. \label{eq01}
\end{equation}
Therefore the main superfield of the theory must have an odd number of indices. Hence
the construction of a higher integer superspin theory must be developed around a spinorial
superfield $\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}$.
\section{The Setup}
~~As our starting point, we assume the main object of the theory to be a spinorial superfield
$\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}$, this means it's highest spin component (the $\theta\bar{\theta}$ term)
must be a propagating fermion. Therefore the mass dimensions of $\Psi$ must be $1/2$.
The most general action which is quadratic for a spinorial superfield $\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}$ with
mass dimensions $\[\Psi\]=1/2$ has to include exactly 2 ${\rm D}$'s (${\bar{\rm D}}$'s) and thus takes the
form:
\begin{equation}
\eqalign{
S=\int d^8z \Big\{& \,
c_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
&+c_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
&+\,a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar{\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
&+a_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\Big\} ~~.
} \label{eq02}
\end{equation}
Since this action is going to describe a massless supermultiplet, it should be invariant under
a gauge symmetry. The most general gauge symmetry allowed by the invariance of the higher
spin projection operator \eqref{eq01} has the following structure:
\begin{equation}
\delta\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}=\[\frac{1}{s!}\]{\rm D}_{({\alpha}_s}K_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}+\[\frac{1}{(s-1)!}\]{\rm D}_{
({\dot{\alpha}}_{s-1}}\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2))} ~~. \label{eq03}
\end{equation}
The change of the above action with respect to the gauge transformation is:
\begin{equation}
\eqalign{
\delta S=\int d^8z \Bigg\{&\left(-2c_1{\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+a_2{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi
}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right){\rm D}^{{\beta}}{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
&+2c_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2{\rm D}_{{\alpha}_s}K_{{\alpha}(s-1){\dot{\alpha}}(s-1)}
-a_1\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}D^2{\bar{\rm D}}_{{\dot{\alpha}}_s}K_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
&+\left(a_2\[\frac{s+1}{s}\]-a_1\right)\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2K_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
&+a_1\[\frac{s-1}{s}\]\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}{\rm D}_{{\alpha}_{s-1}}{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^{{\beta}}K_{{\beta}{\alpha}(s-2){\dot{\alpha}}(s-1)}
+c.c.\Bigg\} ~~.
}
\label{eq04}
\end{equation}
This expression will play a key role in discovering the different ways an integer higher spin
superspin theory can be realized. Our ultimate goal is, based on the above equations
\eqref{eq03} and \eqref{eq04}, to find all possible ways to build a gauge invariant action which
on-shell has exactly the degrees of freedom to form a massless irreducible representation
of the super Poincare group.
Following this path suggests the special case of $s=1$ has to be treated
separately since the index structure drastically changes\footnote{The structure of the $\Lambda$
term in the gauge transformation \eqref{eq03} has to change and equation \eqref{eq04}
\\ $~~~~~~$ is
simplified considerably.}. At this point we focus on the $s>1$ case
\section{Consideration for ${\bm s>1}$ Case}
~~~Now all we have to do is to find and introduce a set of appropiate and unconstraint
compensators that will serve a double purpose. First of all they must give rise to a gauge
invariant action and secondly this action on-shell must generate an irreducible representation
of the super Poincare group. The second requirement, which is very restrictive, can be phrased
in a different way. We can interpret it to say the invariant action constructed out of the superfield $\Psi$
and a set of compensators when expanded in component fields must give the massless
integer spin Fronsdal action for its bosonic piece and the massless half integer Fronsdal
action for its fermionic piece. Since $\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}$ doesn't include all the fields
needed for the Fronsdal actions, we need at least one propagating compensator. In
principle it can be either a bosonic superfield\footnote{This must possess a mass dimensions
0.} or a fermionic superfield\footnote{This possesses mass dimensions 1/2.}.
Any attempt to introduce a fermionic compensator which has less indices than the main
superfield appears precluded by the gauge invariance of the action. In order to succeed
with the fermionic compensator it must be constrained\footnote{The constraint can be
solved in terms of an unconstrained bosonic superfield.}. Therefore the propagating
compensator needed has to be a bosonic one $V$ with zero mass dimensions. This
follows the pattern of the higher spin theories developed so far in that the statistics flip
between the main superfield and the propagating prepotential
One more very important observation is that $V$, must have a gauge transformation
that involves either $K$ or $\Lambda$ parameters, since these are the only two available.
But these parameters also have mass dimensions 0. That means that the gauge
transformation will be algebraic (no derivatives are present). Ignoring the index structure
and coefficients this gauge variation must look as
\begin{equation}
\delta V \sim K + \Lambda ~~~. \nonumber
\end{equation}
~~~This is not acceptable because it means that $V$ can be {\em {completely}} gauged away
and therefore there is nothing around to provide the extra degrees of freedom needed in order
to form on-shell an irreducible multiplet. The only way out is if we allow the gauge parameters
$K$ or $\Lambda$ to have some more ${\rm D}$-structure within them. In this way we could
introduce a zero mass dimensions bosonic compensator with a gauge transformation
which is not algebraic and therefore could be used to gauge away all the unwanted
degrees of freedom.
The last piece of information that we need in order to construct the full theory, is what type
of gauge transformations for $V$, should we introduce? The structure of the Fronsdal action
comes to the rescue. In the massless integer spin Fronsdal action, there are two real
bosonic component fields:
\begin{equation}
\eqalign{
&\text{the main field}~~ h_{{\alpha}(s){\dot{\alpha}}(s)},~~~~\[h_{{\alpha}(s){\dot{\alpha}}(s)}\]=1,~~~\delta h_{{\alpha}(s){\dot{\alpha}}(s)}=
\[\frac{1}{s!^2}\]\pa_{({\alpha}_s({\dot{\alpha}}_s}\zeta_{{\alpha}(s-1)){\dot{\alpha}}(s-1))}~~, \cr \nonumber
&\text{a compensator}~ h_{{\alpha}(s-2){\dot{\alpha}}(s-2)},~\[h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\]=1,~~\delta
h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}=\pa^{{\dot{\alpha}}_s{\alpha}_s}\zeta_{{\alpha}(s-1){\dot{\alpha}}(s-1)} ~.
}
\end{equation}
The main superfield $\Psi$ can provide a component field with the index structure of $h_{{\alpha}(s)
{\dot{\alpha}}(s)}$, but not one for the role of $h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}$. This field has to come from the
compensator $V$. So on-shell $V$ must provide only one real bosonic component with
the proper index structure, mass dimensions and gauge transformation in order to play
the role of $h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}$. This suggest that:
\begin{itemize}
\item $V$ should be real and therefore it's index structure must be $V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$,
\item The $V^{(0,0)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ component must be able to be gauged away, it
has wrong mass dimensions and index structure. This can be achieved if:
\begin{equation}
\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}| \sim \text{~some component of the gauge parameter
(algebraically)},
\nonumber
\end{equation}
\item The $V^{(2,0)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ must be able to be gauged away (wrong index
structure)
\begin{equation}
{\rm D}^2 \delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}|\sim \text{~some component of the gauge parameter
(algebraically),}\nonumber
\end{equation}
\item The $V^{(1,1)(S,S)}_{{\alpha}(s){\dot{\alpha}}(s)}$ component must be able to be gauged away\\
(wrong index structure)
\begin{equation} \eqalign{
\[{\rm D}_{({\alpha}_s},{\bar{\rm D}}_{({\dot{\alpha}}_s}\]\delta V_{{\alpha}(s-1)){\dot{\alpha}}(s-1))}| &\sim \text{~some component of
the gauge parameter } \cr
& \text{~~~~ (algebraically), ~and}
\nonumber
} \end{equation}
\item The $V^{(1,1)(A,A)}_{{\alpha}(s-2){\dot{\alpha}}(s-2)}$ component must survive on-shell and
transform like
\begin{equation} \eqalign{
\[{\rm D}^{{\alpha}_{s-1}},{\bar{\rm D}}^{({\dot{\alpha}}_{s-1}}\]\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}|&\sim \pa^{{\dot{\alpha}}_{s-1}{\alpha}_{s-1}}\text{~
some component of the gauge} \cr
& \text{~~~~\,parameter.} \nonumber
} \end{equation}
\end{itemize}
These requirements fix the desired gauge transformation for the bosonic compensator
to the following form:
\begin{equation}
\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}={\rm D}^{{\alpha}_s}U_{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^{{\dot{\alpha}}_s}\bar{U}_{{\alpha}(s-1){\dot{\alpha}}(s)}
~~.
\label{eq05}
\end{equation}
Now it is very clear for what we are searching. By studying equation \eqref{eq04}, we
will explore all possible ways that we can introduce a real bosonic compensator with
the above gauge transformation \eqref{eq05}. This can happen only by a couple of ways.
The first thing that is clear is that the parameter $\Lambda$, because of it's index structure
and the way it appears in \eqref{eq04}, can not have an internal structure such it that will
lead to the introduction of the desired compensator. So our efforts must focus on the
parameter $K$. That also means that if we insist on having a $\Lambda$-term in \eqref{eq03}
we must introduce another compensator which must be auxiliary\footnote{in order not to
introduce new degrees of freedom} or the coefficients related to the $\Lambda$ terms must
vanish.
\subsection{The KS-series}
~~~~By observing \eqref{eq04} we see that if $K_{{\alpha}(s-1){\dot{\alpha}}{s-1}}={\rm D}^{{\alpha}_s}U_{{\alpha}(s){\dot{\alpha}}
(s-1)}$ then the last two terms vanish and the change of the action becomes
\begin{equation}
\eqalign{
\delta S=\int d^8z &\left(-2c_1{\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+a_2{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}
(s-1){\dot{\alpha}}(s)}\right){\rm D}^{{\beta}}{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
&+\left(2c_2{\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}-a_1{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1)
{\dot{\alpha}}(s)}\right){\rm D}^{{\beta}}U_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
&+c.c.
} \label{equ06}
\end{equation}
Now it is obvious that if
\begin{equation}
\eqalign{
-2c_1&=a_2 ~~,~~
2c_2=-a_1 ~~~,
} \label{equ07}
\end{equation}
the variation in (\ref{equ06}) becomes
\begin{equation}
\eqalign{
\delta S=\int d^8z\ &a_2\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1
){\dot{\alpha}}(s)}\right)\left[{\rm D}^{{\beta}}{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2)}+c.c.\right]\cr
-&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1)
{\dot{\alpha}}(s)}\right)\left[{\rm D}^{{\beta}}U_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\right]
~~. } \label{equ08}
\end{equation}
The observations above suggest the introduction of two real compensators\\
$B_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ with mass dimensions $[B]=1,~[V]=0$ and
gauge transformations
\begin{equation}
\eqalign{
&\delta B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=\[\frac{1}{(s-1)!}\]\Big({\rm D}^{{\alpha}_s}{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}\Lambda_{
{\alpha}(s){\dot{\alpha}}(s-2))}+{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{({\alpha}_{s-1}}\bar{\Lambda}_{{\alpha}(s-2)){\dot{\alpha}}(s)}\Big) ~~, \cr
&\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}={\rm D}^{{\alpha}_s}U_{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^{{\dot{\alpha}}_s}\bar{U}_{{\alpha}(s-1){\dot{\alpha}}(s)}
~~. } \label{equ09}
\end{equation}
The gauge transformation of $\Psi$ becomes
\begin{equation}
\delta\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}=-{\rm D}^2U_{{\alpha}(s){\dot{\alpha}}(s-1)}+\[\frac{1}{(s-1)!}\]{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}
\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2))} ~~. \label{equ10}
\end{equation}
In order to construct a gauge invariant action and the compensators to have dynamics
we add the following terms in the action:
\begin{itemize}
\item Counter terms (they cancel the change of the initial action)
\begin{equation}
\eqalign{
S_{c}=\int d^8z \Bigg\{-&a_2\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{
{\alpha}(s-1){\dot{\alpha}}(s)}\right)B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right)
V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Bigg\} ~~,
} \label{equ11}
\end{equation}
\item Kinetic energy terms (the most general free action for each of the compensators)
\begin{equation}
\eqalign{
S_{k.e}=\int d^8z~&eB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_1 V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}} V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_2 V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_3V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}\pa^{{\dot{\gamma}}{\gamma}} V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(
s-2)}\cr
+&h_4 V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\right]\left[{\rm D}^{{\gamma}},
{\bar{\rm D}}^{{\dot{\gamma}}}\right] V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)} ~~,
} \label{equ12}
\end{equation}
\item Interaction terms (in principle there might be interactions among compensators)
\begin{equation}
\eqalign{
S_{int.}=\int d^8z bB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1){\dot{\alpha}}(
s-1)}\right) ~~. } \label{equ13}
\end{equation}
\end{itemize}
Thus the full action look as below and contains only a series of constants to be determined.
\begin{equation}
\eqalign{ {~~~~~}
S=\int d^8z\Bigg\{
-&\frac{1}{2}a_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&\,a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&a_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
-&a_2\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right)
B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1)
{\dot{\alpha}}(s)}\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&eB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_2V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_3V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}\pa^{{\dot{\gamma}}{\gamma}}V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&h_4V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\right]\left[{\rm D}^{{\gamma}},{\bar{\rm D}}^{{\dot{\gamma}}}
\right]V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&bB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\right)\Bigg\} ~~.
} \label{equ14}
\end{equation}
The requirement that this action is invariant under the above transformations, will give
rise to two Bianchi identities required for the gauge invariance of the action.
\begin{equation}
\eqalign{
&{\rm D}^2{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)}+\[\frac{1}{s!}\]{\rm D}_{({\alpha}_s}{\bm{\cal P}}_{{\alpha}(s-1))
{\dot{\alpha}}(s-1)}=0 ~~, \cr
&{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)}-\[\frac{1}{s!}\]{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}D_{
({\alpha}_s}{\bm{\cal G}}_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}=0 ~~,
} \label{equ15}
\end{equation}
where ${\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)},~{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~{\bm{\cal G}}_{{\alpha}(s-1)
{\dot{\alpha}}(s-1)}$ are the variations of the action with respect to the superfields $\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)
},~V_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$. The solution of the first one gives:
\begin{equation}
\eqalign{
&h_1=\frac{1}{2}a_1 ~~~,~~~ h_2=0~~~,~~
h_3=0 ~~~,~~~h_4=0 ~~~,~~~
b=0 ~~~,
} \label{equ16}
\end{equation}
and the second one gives:
\begin{equation}
\eqalign{
&e=-\frac{1}{2}a_2 ~~~,~~~ b=0 ~~~. \cr
} \label{equ17}
\end{equation}
So the gauge invariant action is:
\begin{equation}
\eqalign{
S=\int d^8z\Bigg\{
-&\frac{1}{2}a_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&\,a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&a_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
-&a_2\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}
\right)B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{
{\alpha}(s-1){\dot{\alpha}}(s)}\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
-&\frac{1}{2}a_2B^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\frac{1}{2}a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}
\Bigg\} ~~~.
} \label{equ18}
\end{equation}
Now we can integrate out the auxiliary superfield $B$. Using the on-shell equation of
motion of $B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ and substitute it back in to the action we get:
\begin{equation}
\eqalign{
S=\int d^8z\Bigg\{
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&\,a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar {\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right)
V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\frac{1}{2}a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Bigg\}
~~, } \label{equ19}
\end{equation}
and this action is invariant under the transformations
\begin{equation}
\eqalign{
&\delta\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}=-{\rm D}^2U_{{\alpha}(s){\dot{\alpha}}(s-1)}+\[\frac{1}{(s-1)!}\]{\bar{\rm D}}_{({\dot{\alpha}}_{s-1
}}\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2))}\cr
&\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}={\rm D}^{{\alpha}_s}U_{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^{{\dot{\alpha}}_s}\bar{U}_{{\alpha}(s-1){\dot{\alpha}}(s)}
~~. } \label{equ20}
\end{equation}
This theory\footnote{We could have reached the same result if from the very begging we
have choosen \\ $~~~~~~$ $c_1=a_2=0$ instead of introducing the auxiliary compensator
$B$.} is equivalent to the theory of S. Kuzenko and A. Sibiryakov \cite{Off2}, once one solves
the constraints that appear in their description (as done in \cite{Gates:2010td}).
This theory is well studied and it is known to describe on-shell a massless supermultiplet of superspin $Y$=$s$.
\subsection{The FVdWH-series}
~~~~Again by observing equation \eqref{eq04} we find that there is another way to arrange
things. By setting $K_{{\alpha}(s-1){\dot{\alpha}}(s-1)}={\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}$ then we find:
\begin{equation}
\eqalign{
\delta S=\int d^8z &\left(-2c_1{\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+a_2{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{
{\alpha}(s-1){\dot{\alpha}}(s)}\right){\rm D}^{{\beta}}{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
&+2c_2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}
-a_1\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}D^2{\bar{\rm D}}_{{\dot{\alpha}}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
&+\left(a_2\[\frac{s+1}{s}\]-a_1\right)\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2
{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
&+a_1\[\frac{s-1}{s}\]\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}{\rm D}_{{\alpha}_{s-1}}{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^{{\beta}}
{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\beta}{\alpha}(s-2){\dot{\alpha}}(s)}\cr
&+c.c.
} \label{equ21}
\end{equation}
and this suggest setting
\begin{equation}
a_2\[\frac{s+1}{s}\]=a_1 ~~, \label{equ22}
\end{equation}
so that we find
\begin{equation}
\eqalign{
\delta S=\int d^8z &\left(-2c_1{\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+a_2{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{
\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right){\rm D}^{{\beta}}{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
&+\left(2c_2{\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}-a_1{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{
{\alpha}(s-1){\dot{\alpha}}(s)}\right){\bar{\rm D}}^{{\dot{\beta}}}U_{{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-1)}\cr
&-a_1\[\frac{s-1}{s}\]{\rm D}^{{\alpha}_{s-1}}{\bar{\rm D}}_{{\dot{\beta}}}{\rm D}_{{\beta}}\bar{\Psi}^{{\beta}{\alpha}(s-2){\dot{\beta}}{\dot{\alpha}}(s-1)}\left(
{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}+c.c.\right)\cr
&+c.c.
} \label{equ23}
\end{equation}
In order to minimise the degrees of freedom that we have to introduce and construct
a minimal theory we set:
\begin{equation}
\eqalign{
-2c_1&=a_2 ~~,~~
2c_2=-a_1 ~~,
} \label{equ24}
\end{equation}
so the change of the action takes the form
\begin{equation}
\eqalign{
\delta S=\int d^8z ~&a_2{\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}\[\frac{1}{(s-1)!}\]\[{\rm D}^{{\beta}}
{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2))}+c.c.\]\cr
+&a_2{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\[\frac{1}{(s-1)!}\]\left[{\rm D}^{{\beta}}{\bar{\rm D}}_{
({\dot{\alpha}}_{s-1}}\Lambda_{{\beta}{\alpha}(s-1){\dot{\alpha}}(s-2))}+c.c.\right]\cr
-&a_1{\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}\left[{\bar{\rm D}}^{{\dot{\beta}}}U_{{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-1)}
+c.c.\right]\cr
-&a_1{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\left[{\bar{\rm D}}^{{\dot{\beta}}}U_{{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}
(s-1)}+c.c.\right]\cr
-&a_1\[\frac{s-1}{s}\]\left({\rm D}^{{\alpha}_{s-1}}{\bar{\rm D}}_{{\dot{\beta}}}{\rm D}_{{\beta}}\bar{\Psi}^{{\beta}{\alpha}(s-2
){\dot{\beta}}{\dot{\alpha}}(s-1)}+c.c.\right)\left[{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}+c.c.\right] ~~~.
} \label{equ25}
\end{equation}
We introduce two real compensators, $B_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$
with $[B]=1,~[V]=0$ and gauge transformations
\begin{equation}
\eqalign{
&\delta B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=\[\frac{1}{(s-1)!}\]\left[{\rm D}^{{\alpha}_s}{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}
\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2))}+{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{({\alpha}_{s-1}}\bar{\Lambda}_{{\alpha}(s-2
)){\dot{\alpha}}(s)}\right] ~~, \cr
&\delta V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}={\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1){\dot{\alpha}}(s)}+{\rm D}^{{\alpha}_s}\bar{
U}_{{\alpha}(s){\dot{\alpha}}(s-1)} ~~,
} \label{equ26}
\end{equation}
and the gauge transformation of the $\Psi$ superfield is
\begin{equation}
\delta\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}=\[\frac{1}{s!}\]{\rm D}_{({\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}U_{{\alpha}(s-1))
{\dot{\alpha}}(s)}+\[\frac{1}{(s-1)!}\]{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2))} ~~~.
\label{equ27}
\end{equation}
Hence we have to add a few terms to the action
\begin{itemize}
\item Counter terms (they cancel the change of the initial action)
\begin{equation}
\eqalign{
S_c=\int d^8z\Big\{-&\[\frac{s}{s+1}\]a_1\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}
\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\right)B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1){\dot{\alpha}}(s)
}\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\[\frac{s-1}{s}\]\left({\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\rm D}_{{\beta}}{\bar{\rm D}}_{{\dot{\beta}}}\Psi^{{\beta}{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-2)}+
c.c.\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)} ~~,
} \label{equ28}
\end{equation}
\item Kinetic energy (the most general action for each of the compensators)
\begin{equation}
\eqalign{
S_{k.e}=\int d^8z
~&eB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_2V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_3V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}\pa^{{\gamma}{\dot{\gamma}}}V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&h_4V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\right]\left[{\rm D}^{{\gamma}},{\bar{\rm D}}^{{\dot{\gamma}}}
\right]V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)} ~~,
} \label{equ29}
\end{equation}
\item Interaction terms (in principle there might be interactions among compensators)
\begin{equation}
S_{int}=\int d^8z bB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1)
{\dot{\alpha}}(s-1)}\right) ~~. \label{equ30}
\end{equation}
\end{itemize}
Therefore the full action is
\begin{equation}
\eqalign{ {~~~~~}
S=\int d^8z \Big\{
-&\frac{1}{2}\[\frac{s}{s+1}\]a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar{\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&\[\frac{s}{s+1}\]a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\Bar {\Psi}}_{{\alpha}(s-1
){\dot{\alpha}}(s)}\cr
-&\[\frac{s}{s+1}\]a_1\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1
){\dot{\alpha}}(s)}\right)B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1)
{\dot{\alpha}}(s)}\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\[\frac{s-1}{s}\]a_1\left({\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\rm D}_{{\beta}}{\bar{\rm D}}_{{\dot{\beta}}}\Psi^{{\beta}{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-2
)}+c.c.\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&eB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_2V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&h_3V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}\pa^{{\gamma}{\dot{\gamma}}}V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&h_4V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1}}\right]\left[{\rm D}^{{\gamma}},{\bar{\rm D}}^{{\dot{\gamma}}}
\right]V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&bB^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\right)\Big\}
~~~. } \label{equ31}
\end{equation}
The invariance of this action under the gauge transformations is guaranteed by the
following two Bianchi Identities (derived as before):
\begin{equation}
\eqalign{
\[\frac{1}{s!}\]{\bar{\rm D}}_{({\dot{\alpha}}_s}{\rm D}^{{\alpha}_s}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1))}-\[\frac{1}{s!}\]{\bar{\rm D}}_{({\dot{\alpha}}_s}
{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1))}=0 ~~, \cr
{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)}-\[\frac{1}{s!}\]{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\rm D}_{({\alpha}_s}{\bm{\cal
G}}_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}=0 ~~,
} \label{equ32}
\end{equation}
where ${\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)},~{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~{\bm{\cal G}}_{{\alpha}
(s-1){\dot{\alpha}}(s-1)}$ are the variations of the action with respect to the superfields $\Psi_{
{\alpha}(s){\dot{\alpha}}(s-1)},~V_{{\alpha}(s-1){\dot{\alpha}}(s-1)},~B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$.
The solution of the first one gives:
\begin{equation}
\eqalign{
&h_1=\[\frac{1}{2s}\]a_1 ~~~,~~~~~~~~~~~~\,~~~~~~~ h_2=-\[\frac{s-1}{2s}\]a_1 ~~,\cr
&h_3=\[\frac{(2s-1)(s-1)}{(2s)^2}\]a_1 ~~~,~~~ h_4=\[\frac{s-1}{(2s)^2}\]a_1 ~,\cr
&b=-a_1 ~~~,
} \label{equ33}
\end{equation}
and the second one has as a solution:
\begin{equation}
\eqalign{
&e=-\frac{1}{2}\[\frac{s}{s+1}\]a_1 ~~~,~~~ b=-a_1 ~~.
} \label{equ34}
\end{equation}
Thus the action takes the form
\begin{equation}
\eqalign{ {~~~~~}
S=\int d^8z \Big\{
-&\frac{1}{2}\[\frac{s}{s+1}\]a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar{\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&\[\frac{s}{s+1}\]a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\rm D}_{{\alpha}_s}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\Bar {\Psi}}_{{\alpha}(s-1)
{\dot{\alpha}}(s)}\cr
-&\[\frac{s}{s+1}\]a_1\left({\rm D}_{{\alpha}_s}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}\bar{\Psi}^{{\alpha}(s-1
){\dot{\alpha}}(s)}\right)B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\left({\rm D}_{{\alpha}_s}{\bar{\rm D}}^2\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}_{{\dot{\alpha}}_s}{\rm D}^2\bar{\Psi}^{{\alpha}(s-1)
{\dot{\alpha}}(s)}\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\[\frac{s-1}{s}\]a_1\left({\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\rm D}_{{\beta}}{\bar{\rm D}}_{{\dot{\beta}}}\Psi^{{\beta}{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-2)}
+c.c.\right)V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
-&\frac{1}{2}\[\frac{s}{s+1}\]a_1B^{{\alpha}(s-1){\dot{\alpha}}(s-1)}B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\frac{1}{2s}a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
-&\[\frac{s-1}{2s}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\[\frac{(2s-1)(s-1)}{(2s)^2}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}
\pa^{{\gamma}{\dot{\gamma}}}V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&\[\frac{s-1}{(2s)^2}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1
}}\right]\left[{\rm D}^{{\gamma}},{\bar{\rm D}}^{{\dot{\gamma}}}\right]V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
-&a_1B^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1
)}\right)\Big\} ~~~.
} \label{equ35}
\end{equation}
At this point we can integrate out the auxiliary superfield $B$ by using it's equation
of motion
\begin{equation}
\eqalign{
B_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=~&\left[{\rm D}^{{\alpha}_s}\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^{{\dot{\alpha}}_s}\bar{\Psi}_{{\alpha}(s-1
){\dot{\alpha}}(s)}\right]\cr
-&\[\frac{s+1}{s}\]\left({\rm D}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+{\bar{\rm D}}^2V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\right) ~~,
} \label{equ36}
\end{equation}
and substituting this result back in (\ref{equ35}) yields
\begin{equation}
\eqalign{
S=\int d^8z \Big\{
-&\frac{1}{2}a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)} +c.c. \cr
+&a_1\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{{\alpha}_s}{\Bar{\Psi}}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&a_1\left\{{\bar{\rm D}}^2,{\rm D}_{{\alpha}_s}\right\}\Psi^{{\alpha}(s){\dot{\alpha}}(s-1)}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\cr
+&\[\frac{s-1}{s}\]a_1{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\rm D}_{{\beta}}{\bar{\rm D}}_{{\dot{\beta}}}\Psi^{{\beta}{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-2)}
V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\cr
+&\[\frac{s+2}{2s}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\[\frac{1}{s}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&\[\frac{(2s-1)(s-1)}{(2s)^2}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}
\pa^{{\gamma}{\dot{\gamma}}}V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\cr
+&\[\frac{s-1}{(2s)^2}\]a_1V^{{\alpha}(s-1){\dot{\alpha}}(s-1)}\left[{\rm D}_{{\alpha}_{s-1}},{\bar{\rm D}}_{{\dot{\alpha}}_{s-1
}}\right]\left[{\rm D}^{{\gamma}},{\bar{\rm D}}^{{\dot{\gamma}}}\right]V_{{\gamma}{\alpha}(s-2){\dot{\gamma}}{\dot{\alpha}}(s-2)}\Big\} ~~. \label{action}
}
\end{equation}
Calculating variations with respect to $\Psi$ and $V$ in this action we can define the following superfields
\begin{equation}
\eqalign{
{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)}=-&a_1{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+\frac{a_1}{s!}
{\bar{\rm D}}^{{\dot{\alpha}}_s}{\rm D}_{({\alpha}_s}\bar{\Psi}_{{\alpha}(s-1)){\dot{\alpha}}(s)} {~~~~~~~~~~~~~~} \cr
+&\frac{a_1}{s!}{\bar{\rm D}}^2{\rm D}_{({\alpha}_s}V_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}+\frac{a_1}{s!}{\rm D}_{({\alpha}_s}
{\bar{\rm D}}^2V_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}\cr
-&\frac{s-1}{s!s!}a_1{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}{\rm D}_{({\alpha}_s}{\bar{\rm D}}^{{\dot{\beta}}}V_{{\alpha}(s-1)){\dot{\beta}}{\dot{\alpha}}(s-2))}
~~, } \label{equ38}
\end{equation}
$$
\eqalign{ {~~~~~}
{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=-&a_1{\rm D}^{{\alpha}_s}{\bar{\rm D}}^2\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}-a_1{\bar{\rm D}}^{{\dot{\alpha}}_s}
{\rm D}^2\bar{\Psi}_{{\alpha}(s-1){\dot{\alpha}}(s)} {~~~~~~~~~~~~~~~~~~~~} \cr
-&a_1{\bar{\rm D}}^2{\rm D}^{{\alpha}_s}\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}-a_1{\rm D}^2{\bar{\rm D}}^{{\dot{\alpha}}_s}\bar{\Psi}_{{\alpha}(s-1){\dot{\alpha}}(s)}\cr
+&a_1\frac{s-1}{s!}{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}{\rm D}^{{\beta}}{\bar{\rm D}}^{{\dot{\beta}}}\Psi_{{\beta}{\alpha}(s-1){\dot{\beta}}{\dot{\alpha}}(s-2))}\cr
+&a_1\frac{s-1}{s!}{\rm D}_{({\alpha}_{s-1}}{\bar{\rm D}}^{{\dot{\beta}}}{\rm D}^{{\beta}}\bar{\Psi}_{{\beta}{\alpha}(s-2)){\dot{\beta}}{\dot{\alpha}}(s-1)}\cr
+&a_1\frac{s+2}{s}{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
+&a_1\frac{2}{s}\Box V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\cr
}
$$
\begin{equation}
\eqalign{ {~~~~~}
~~~~~~~~~~~~~~+&a_1\frac{(2s-1)(s-1)}{2s!^2}\pa_{({\alpha}_{s-1}({\dot{\alpha}}_{s-1}}\pa^{{\beta}{\dot{\beta}}}
V_{{\beta}{\alpha}(s-2)) {\dot{\beta}}{\dot{\alpha}}(s-2))}\cr
+&a_1\frac{s-1}{2s!^2}\left[{\rm D}_{({\alpha}_{s-1}},{\bar{\rm D}}_{({\dot{\alpha}}_{s-1}}\right]\left[{\rm D}^{{\beta}},
{\bar{\rm D}}^{{\dot{\beta}}}\right]V_{{\beta}{\alpha}(s-2)){\dot{\beta}}{\dot{\alpha}}(s-2))} ~~,
} \label{equ39}
\end{equation}
and they satisfy the Bianchi Identities for this final action
\begin{equation}
\eqalign{
&{\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1)}=0 ~~, \cr
&\frac{1}{s!}{\bar{\rm D}}_{({\dot{\alpha}}_s}{\rm D}^{{\alpha}_s}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1))}-\frac{1}{s!}{\bar{\rm D}}_{({\dot{\alpha}}_s}
{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1))}=0 ~~.
} \label{equ40}
\end{equation}
Furthermore we can prove that they also satisfy the following identity:
\begin{equation}
\eqalign{
{\bar{\rm D}}^{{\dot{\alpha}}_{2s}}\bar{\bm{\cal W}}_{{\dot{\alpha}}(2s)}=&-i\frac{2s}{a_1}\pa^{{\alpha}_{s}}{}_{({\dot{\alpha}}_{2s-1
}}\dots\pa^{{\alpha}_{1}}{}_{{\dot{\alpha}}_{s}}{\bm{\cal T}}_{{\alpha}(s){\dot{\alpha}}(s-1))}\cr
&-\frac{s}{a_1}{\bar{\rm D}}^2\pa^{{\alpha}_{s-1}}{}_{({\dot{\alpha}}_{2s-1}}\dots\pa^{{\alpha}_{1}}{}_{{\dot{\alpha}}_{s+1}}
\Bar{\bm{\cal T}}_{{\alpha}(s-1){\dot{\alpha}}(s))}\cr
&+\frac{s}{a_1}{\bar{\rm D}}_{({\dot{\alpha}}_{2s-1}}\pa^{{\alpha}_{s-1}}{}_{{\dot{\alpha}}_{2s-2}}\dots\pa^{{\alpha}_{1}}
{}_{{\dot{\alpha}}_{s}}{\bm{\cal P}}_{{\alpha}(s-1){\dot{\alpha}}(s-1))} ~~,
} \label{equ42}
\end{equation}
where the anti-chiral superfield (i.e. ${\rm D}_{{\beta}} \bar{\bm{\cal W}}_{{\dot{\alpha}}(2s)}= 0$) is given by
\begin{equation}
\bar{\bm{\cal W}}_{{\dot{\alpha}}(2s)}={\rm D}^2{\bar{\rm D}}_{({\dot{\alpha}}_{2s}}\pa^{{\alpha}_{s-1}}{}_{{\dot{\alpha}}_{2s-1}}\dots
\pa^{{\alpha}_{1}}{}_{{\dot{\alpha}}_{s+1}}\bar{\Psi}_{{\alpha}(s-1){\dot{\alpha}}(s))} ~~. \label{equ42}
\end{equation}
So this theory has an irreducible multiplet propagating on-shell with superspin $Y=s$.
Now we can check if these are the only degrees of freedom propagating.\\
Expanding the superfields $\Psi,~V$ into components and using their gauge transformations
we find that some components have purely algebraic transformations and therefore can
be gauged away. In detail\footnote{The definition of symmetric and antisymmetric
pieces of a field is the following $$\Phi_{{\gamma}{\alpha}(s-1)}=\Phi^{(S)}_{{\gamma}{\alpha}(s-1)}+\frac{s-1}{s!}
C_{{\gamma}({\alpha}_{s-1}}\Phi^{(A)}_{{\alpha}(s-2))}~~,~~\Phi^{(S)}_{{\gamma}{\alpha}(s-1)}=\frac{1}{s!}\Phi_{({\gamma}{\alpha}(s-1))}
~~,~~\Phi^{(A)}_{{\alpha}(s-2)}=C^{{\gamma}{\alpha}_{s-1}}\Phi_{{\gamma}{\alpha}(s-1)}$$
Furthermore the notation $\Phi^{(m,n)}$ represents the $\theta^{m}\bar{\theta}^n$
component in the taylor series of the superfield $\Phi$}:\vskip.2in
\begin{tabular}{c c}
For Bosons & {~~} For Fermions
\\
\begin{tabular}{| l | l |}
\hline
Component & Gauged away by\\
\hline
$V^{(0,0)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $Re\left[U^{(0,1)(A)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}\right]$\\
\hline
$V^{(1,1)(S,S)}_{{\alpha}(s){\dot{\alpha}}(s)}$ & $Re\left[U^{(1,2)(S)}_{{\alpha}(s){\dot{\alpha}}(s)}\right]$\\
\hline
$V^{(1,1)(A,S)}_{{\alpha}(s-2){\dot{\alpha}}(s)}$ & $U^{(1,2)(A)}_{{\alpha}(s-2){\dot{\alpha}}(s)}$\\
\hline
$V^{(2,0)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $U^{(2,1)(A)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$\\
\hline
$\Psi^{(1,0)(S)}_{{\alpha}(s+1){\dot{\alpha}}(s-1)}$ & $\Lambda^{(1,1)(S,S)}_{{\alpha}(s+1){\dot{\alpha}}(s-1)}$\\
\hline
$\Psi^{(1,0)(A)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $\Lambda^{(1,1)(A,S)}_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$\\
\hline
$Im\left[\Psi^{(0,1)(S)}_{{\alpha}(s){\dot{\alpha}}(s)}\right]$ & $Im\left[U^{(1,2)(S)}_{{\alpha}(s){\dot{\alpha}}(s)}\right]$\\
\hline
$\Psi^{(2,1)(A)}_{{\alpha}(s){\dot{\alpha}}(s-2)}$ & $\Lambda^{(2,2)}_{{\alpha}(s){\dot{\alpha}}(s-2)}$\\
\hline
\end{tabular}
& {~~~~}
\begin{tabular}{| l | l |}
\hline
Component & Gauged away by\\
\hline
$V^{(0,1)(S)}_{{\alpha}(s-1){\dot{\alpha}}(s)}$ & $U^{(0,2)}_{{\alpha}(s-1){\dot{\alpha}}(s)}$\\
\hline
$V^{(1,0)(A)}_{{\alpha}(s-2){\dot{\alpha}}(s-1)}$ & $U^{(1,1)(A,A)}_{{\alpha}(s-2){\dot{\alpha}}(s-1)}$\\
\hline
$V^{(2,1)(S)}_{{\alpha}(s-1){\dot{\alpha}}(s)}$ & $U^{(2,2)}_{{\alpha}(s-1){\dot{\alpha}}(s)}$\\
\hline
$\Psi^{(0,0)}_{{\alpha}(s){\dot{\alpha}}(s-1)}$ & $\Lambda^{(0,1)(S)}_{{\alpha}(s){\dot{\alpha}}(s-1)}$\\
\hline
$\Psi^{(2,0)}_{{\alpha}(s){\dot{\alpha}}(s-1)}$ & $\Lambda^{(2,1)(S)}_{{\alpha}(s){\dot{\alpha}}(s-1)}$\\
\hline
$\Psi^{(1,1)(S,A)}_{{\alpha}(s+1){\dot{\alpha}}(s-2)}$ & $\Lambda^{(1,2)(S)}_{{\alpha}(s+1){\dot{\alpha}}(s-2)}$\\
\hline
$\Psi^{(1,1)(A,A)}_{{\alpha}(s-1){\dot{\alpha}}(s-2)}$ & $\Lambda^{(1,2)(A)}_{{\alpha}(s-1){\dot{\alpha}}(s-2)}$\\
\hline
\end{tabular}
\end{tabular}
So in the Wess-Zumino gauge the two superfields take the forms
\begin{equation}
\eqalign{
V_{{\alpha}(s-1){\dot{\alpha}}(s-1)}&=\[\frac{2(s-1)}{(s-1)!^2}\]\theta_{({\alpha}_{s-1}}\bar{\theta
}_{({\dot{\alpha}}_{s-1}}h_{{\alpha}(s-2)){\dot{\alpha}}(s-2))}\cr
&~~+\[\frac{\sqrt{2}}{(s-1)!}\]\theta^2\bar{\theta}_{({\dot{\alpha}}_{s-1}}\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2))}\cr
&~~-\[\frac{\sqrt{2}}{(s-1)!}\]\theta_{({\alpha}_{s-1}}\bar{\theta}^2\bar{\psi}_{{\alpha}(s-2)){\dot{\alpha}}(s-1)}
+\theta^2\bar{\theta}^2P_{{\alpha}(s-1){\dot{\alpha}}(s-1)} ~~, \label{taylorA}
}
\end{equation}
and
\begin{equation}
\eqalign{
\Psi_{{\alpha}(s){\dot{\alpha}}(s-1)}&=\bar{\theta}^{{\dot{\alpha}}_s}h_{{\alpha}(s){\dot{\alpha}}(s)}+\sqrt{2}\bar{\theta}^2\psi_{
{\alpha}(s){\dot{\alpha}}(s-1)}+\sqrt{2}\theta^{{\alpha}_{s+1}}\bar{\theta}^{{\dot{\alpha}}_{s}}\psi_{{\alpha}(s+1){\dot{\alpha}}(s)}\cr
&~~~+\[\frac{1}{s!}\]\theta_{({\alpha}_s}\bar{\theta}^{{\dot{\alpha}}_{s}}\[\lambda_{{\alpha}(s-1)){\dot{\alpha}}(s)}-\frac{
\sqrt{2}s}{s+1}\bar{\psi}_{{\alpha}(s-1)){\dot{\alpha}}(s)}\]\cr
&~~~+\theta^2\bar{\theta}^{{\dot{\alpha}}_s}Y_{{\alpha}(s){\dot{\alpha}}(s)}\cr
&~~~+\bar{\theta}^2\theta^{{\alpha}_{s+1}}\[t_{{\alpha}(s+1){\dot{\alpha}}(s-1)}+i\frac{3}{2(s+1)!}\pa_{(
{\alpha}_{s+1}}{}^{{\dot{\alpha}}_s}h_{{\alpha}(s)){\dot{\alpha}}(s)}\]\cr
&~~~+\[\frac{s}{(s+1)!}\]\bar{\theta}^2\theta_{({\alpha}_{s}}\[M_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}+\frac{(s+1
)^2}{s(2s+1)}P_{{\alpha}(s-1)){\dot{\alpha}}{s-1}}\]\cr
&~~~+i\[\frac{s}{(s+1)!}\]\bar{\theta}^2\theta_{({\alpha}_{s}}\[N_{{\alpha}(s-1)){\dot{\alpha}}(s-1)}+\frac{2s-1}
{2}\pa^{{\gamma}{\dot{\gamma}}}h_{{\gamma}{\alpha}(s-1)){\dot{\gamma}}{\dot{\alpha}}(s-1)}\right.\cr
&~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~-\left.\frac{(s+1)^2(s-1)}{s!}\pa_{{\alpha}_{s-1}
({\dot{\alpha}}_{s-1}}h_{{\alpha}(s-2)){\dot{\alpha}}(s-2))}\]\cr
&~~~+\theta^2\bar{\theta}^2\[\chi_{{\alpha}(s){\dot{\alpha}}(s-1)}+\frac{i}{\sqrt{2}}\frac{s-1}{s+1}\pa^{
{\alpha}_{s+1}{\dot{\alpha}}_s}\psi_{{\alpha}(s+1){\dot{\alpha}}(s)}\right.\cr
&~~~~-\left.\frac{i}{\sqrt{2}}\frac{s(s-1)}{(s+1)^2s!}\pa_{({\alpha}_s}{}^{{\dot{\alpha}}_s}\bar{\psi}_{{\alpha}(
s-1)){\dot{\alpha}}(s)}-\frac{i}{2(s+1)!}\pa_{({\alpha}_s}{}^{{\dot{\alpha}}_s}\lambda_{{\alpha}(s-1)){\dot{\alpha}}(s)}\]
~~, \label{taylorM}
}
\end{equation}
where the components $t,~M,~N,~P,~Y,~\lambda,~\chi$ will be shown to be auxiliary
fields. All the others are symmetric in all undotted and dotted indices separately and the
components $h_{{\alpha}(s){\dot{\alpha}}(s)},~h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}$ are real. The component
action for all the bosons is:
$$
\eqalign{
S_{Bosons}=\int d^4x \Big\{&-2a_1h^{{\alpha}(s){\dot{\alpha}}(s)}\Box h_{{\alpha}(s){\dot{\alpha}}(s)}
{~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~} \cr
&+sa_1h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_{s}{\dot{\alpha}}_{s}}\pa^{{\dot{\gamma}}{\gamma}}h_{{\gamma}{\alpha}(s-1){\dot{\gamma}}{\dot{\alpha}}(s-1)}\cr
&-\left[2s(s-1)\right]a_1h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_s{\dot{\alpha}}_s}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}
h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\cr
&+\left[2s(2s-1)\right]a_1h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\Box h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\cr
}
$$
\begin{equation}
\eqalign{
{~~~~~~~} &+\left[s(s-2)^2\right]a_1h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\pa_{{\alpha}_{s-2}{\dot{\alpha}}_{s-2}}
\pa^{{\dot{\gamma}}{\gamma}} h_{{\gamma}{\alpha}(s-3){\dot{\gamma}}{\dot{\alpha}}(s-3)}\cr
&+ \frac{1}{2}a_1t^{{\alpha}(s+1){\dot{\alpha}}(s-1)}t_{{\alpha}(s+1){\dot{\alpha}}(s-1)}+c.c.\cr
&+ a_1 \[\frac{2s+1}{s+1}\]M^{{\alpha}(s-1){\dot{\alpha}}(s-1)}M_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\cr
&- a_1 \[\frac{(s+1)^3}{s^2(2s+1)}\]P^{{\alpha}(s-1){\dot{\alpha}}(s-1)}P_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\cr
&+ a_1 \[\frac{1}{s+1}\]N^{{\alpha}(s-1){\dot{\alpha}}(s-1)}N_{{\alpha}(s-1){\dot{\alpha}}(s-1)}+c.c.\cr
&-a_1Y^{{\alpha}(s){\dot{\alpha}}(s)}\bar{Y}_{{\alpha}(s){\dot{\alpha}}(s)} ~~,
} \label{equ45}
\end{equation}
and using the equations of motion for the auxiliary fields
\begin{equation}
\eqalign{
&M_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=0 ~~,~~~
N_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=0 ~~,~~~
P_{{\alpha}(s-1){\dot{\alpha}}(s-1)}=0 ~~~, \cr
&t_{{\alpha}(s+1){\dot{\alpha}}(s-1)}=0 ~~,~~~ Y_{{\alpha}(s){\dot{\alpha}}(s)}=0 ~~~, \cr
} \label{equ46}
\end{equation}
we obtain
\begin{equation}
\eqalign{
S_{Bosons}=\int d^4x \Big\{&-2a_1h^{{\alpha}(s){\dot{\alpha}}(s)}\Box h_{{\alpha}(s){\dot{\alpha}}(s)}
+sa_1h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_{s}{\dot{\alpha}}_{s}}\pa^{{\dot{\gamma}}{\gamma}}h_{{\gamma}{\alpha}(s-1){\dot{\gamma}}{\dot{\alpha}}(s-1)}\cr
&-\left[2s(s-1)\right]a_1h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_s{\dot{\alpha}}_s}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}
h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\cr
&+\left[2s(2s-1)\right]a_1h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\Box h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\cr
&+\left[s(s-2)^2\right]a_1h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\pa_{{\alpha}_{s-2}{\dot{\alpha}}_{s-2}}
\pa^{{\dot{\gamma}}{\gamma}}h_{{\gamma}{\alpha}(s-3){\dot{\gamma}}{\dot{\alpha}}(s-3)} ~~,
} \label{equ47}
\end{equation}
and upon by setting $a_1=-\frac{1}{2}$ we find
\begin{equation}
\eqalign{
S_{Bosons}=\int d^4x \Big\{&h^{{\alpha}(s){\dot{\alpha}}(s)}\Box h_{{\alpha}(s){\dot{\alpha}}(s)}
-\frac{s}{2}h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_{s}{\dot{\alpha}}_{s}}\pa^{{\dot{\gamma}}{\gamma}}h_{{\gamma}{\alpha}(s-1){\dot{\gamma}}{\dot{\alpha}}(s-1)}\cr
&+\left[s(s-1)\right]h^{{\alpha}(s){\dot{\alpha}}(s)}\pa_{{\alpha}_s{\dot{\alpha}}_s}\pa_{{\alpha}_{s-1}{\dot{\alpha}}_{s-1}}h_{{\alpha}(
s-2){\dot{\alpha}}(s-2)}\cr
&-\left[s(2s-1)\right]h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\Box h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}\cr
&-\left[\frac{s(s-2)^2}{2}\right]h^{{\alpha}(s-2){\dot{\alpha}}(s-2)}\pa_{{\alpha}_{s-2}{\dot{\alpha}}_{s-2}}
\pa^{{\dot{\gamma}}{\gamma}}h_{{\gamma}{\alpha}(s-3){\dot{\gamma}}{\dot{\alpha}}(s-3)} ~~~,
} \label{equ48}
\end{equation}
which is the Fronsdal action for a propagating massless spin-$s$ bosonic field.\\
The fermionic piece of the action is:
$$
\eqalign{
S_{Fermions}=\int d^4x\Big\{
&-i2a_1\bar{\psi}^{{\alpha}(s){\dot{\alpha}}(s+1)}\pa^{{\alpha}_{s+1}}{}_{{\dot{\alpha}}_{s+1}}\psi_{{\alpha}(s+1){\dot{\alpha}}(s)}
{~~~~~~~~~~~~~~~~~~~} {~~~~~~~~~~~~~~~~~~} \cr
&+i2\[\frac{2s+1}{(s+1)^2}\]a_1\bar{\psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\pa^{{\alpha}_s}{}_{{\dot{\alpha}}_s}\psi_{
{\alpha}(s){\dot{\alpha}}(s-1)}\cr
&+i2a_1\bar{\psi}^{{\alpha}(s-2){\dot{\alpha}}(s-1)}\pa^{{\alpha}_{s-1}}{}_{{\dot{\alpha}}_{s-1}}\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
}
$$
\begin{equation}
\eqalign{
{~~~\,~} &-i2\[\frac{s}{s+1}\]a_1\psi^{{\alpha}(s+1){\dot{\alpha}}(s)}\pa_{{\alpha}_{s+1}{\dot{\alpha}}_s}\psi_{{\alpha}(s)
{\dot{\alpha}}(s-1)}+c.c.\cr
&+i2a_1\psi^{{\alpha}(s){\dot{\alpha}}(s-1)}\pa_{{\alpha}_s{\dot{\alpha}}_{s-1}}\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2)}+c.c.\cr
&+\[\frac{s+1}{s}\]\bar{\lambda}^{{\alpha}(s){\dot{\alpha}}(s-1)}\chi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.
} \label{equ49}
\end{equation}
so that using the equations of motions for the auxiliary fields $\chi,~\lambda$
\begin{equation}
\lambda_{{\alpha}(s-1){\dot{\alpha}}(s)}=0 ~~, ~~~~ \chi_{{\alpha}(s){\dot{\alpha}}(s-1)}=0 ~~, \label{equ50}
\end{equation}
and setting the value of $a_1=-\frac{1}{2}$,we get the final fermionic action
\begin{equation}
\eqalign{
S_{Fermions}=\int d^4x\Big\{
&i\bar{\psi}^{{\alpha}(s){\dot{\alpha}}(s+1)}\pa^{{\alpha}_{s+1}}{}_{{\dot{\alpha}}_{s+1}}\psi_{{\alpha}(s+1){\dot{\alpha}}(s)}\cr
&-i\[\frac{2s+1}{(s+1)^2}\]\bar{\psi}^{{\alpha}(s-1){\dot{\alpha}}(s)}\pa^{{\alpha}_s}{}_{{\dot{\alpha}}_s}\psi_{{\alpha}(s)
{\dot{\alpha}}(s-1)}\cr
&-i\bar{\psi}^{{\alpha}(s-2){\dot{\alpha}}(s-1)}\pa^{{\alpha}_{s-1}}{}_{{\dot{\alpha}}_{s-1}}\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2)}\cr
&+i\[\frac{s}{s+1}\]\psi^{{\alpha}(s+1){\dot{\alpha}}(s)}\pa_{{\alpha}_{s+1}{\dot{\alpha}}_s}\psi_{{\alpha}(s){\dot{\alpha}}(s-1)}+c.c.\cr
&-i\psi^{{\alpha}(s){\dot{\alpha}}(s-1)}\pa_{{\alpha}_s{\dot{\alpha}}_{s-1}}\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2)}+c.c.
} \label{equ51}
\end{equation}
This is the Frondsal action for a propagating massles spin-$(s+1/2)$
Therefore we conclude that only an irreducible supermultiplet propagates on-shell
and therefore the action \eqref{action} describes a massless integer superspin
$Y$ $=$ $s$.
The counting of the off-shell bosonic and fermionic degrees of freedom for the
action including all the auxiliary fields is:
\begin{center}
\begin{tabular} {| c | c | c |}
\hline
Component Field(s) & Bosonic & Fermionic\\
\hline
$h_{{\alpha}(s){\dot{\alpha}}(s)}$~/~$h_{{\alpha}(s-2){\dot{\alpha}}(s-2)}$ & $s^2 + 2$ & {}\\
\hline
$\psi_{{\alpha}(s+1){\dot{\alpha}}(s)}$~/~$\psi_{{\alpha}(s){\dot{\alpha}}(s-1)}$~/~$\psi_{{\alpha}(s-1){\dot{\alpha}}(s-2)}$ & {} & $4 (
s^2 + s + 1 ) $\\
\hline
$P_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $s^2$ & {}\\
\hline
$\lambda_{{\alpha}(s-1){\dot{\alpha}}(s)}$ & {} & $2 s (s+1) $\\
\hline
$Y_{{\alpha}(s){\dot{\alpha}}(s)} $ & $ 2(s+1)^2$ & {}\\
\hline
$t_{{\alpha}(s+1){\dot{\alpha}}(s-1)} $ & $2(s+2)s$ & {}\\
\hline
$
M_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $s^2$ & {}\\
\hline
$N_{{\alpha}(s-1){\dot{\alpha}}(s-1)}$ & $s^2$ & {}\\
\hline
$\chi_{{\alpha}(s){\dot{\alpha}}(s-1)}$ & {} & $2 s (s+1) $\\
\hline\hline
$~$ & $8s^2+8s+4$ & $8s^2+8s+4$\\
\hline
\end{tabular}
\end{center}
For each case we have verified the existence
of field strength superfields ${\bm{\cal W}}_{{\alpha}(2s)}$, $ {\bm{{\cal P}}}_{{\alpha}(s-1){\dot{\alpha}}(
s-1)}$ and ${\bm{{\cal T}}}_{{\alpha}(s){\dot{\alpha}}(s-1)}$ which occur for both the KS-series and the
FVdWH-series.
\section{Considering the $s = 1$ case}
~~~For the special case of $s=1$, the most general action takes the form:
\begin{equation}
\eqalign{
S=\int d^8z \Big\{
&\, ~a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar{\Psi}}_{{\dot{\alpha}}}
+a_2\Psi^{{\alpha}}{\rm D}_{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\Bar {\Psi}}_{{\dot{\alpha}}} \cr
&~+ \big[ c_1\Psi^{{\alpha}}{\rm D}^2\Psi_{{\alpha}}
~+c_2\Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} +c.c. \big] \Big\} ~~.
} \label{equ52}
\end{equation}
Also in this case, the gauge parameter $\Lambda_{{\alpha}(s){\dot{\alpha}}(s-2)}$, in order to survive,
must be modified to ${\bar{\rm D}}^{{\dot{\alpha}}_{s-1}}\Lambda_{{\alpha}(s){\dot{\alpha}}(s-1)}$. So the most general
gauge transformation allowed in the $s=1$ is:
\begin{equation}
\delta\Psi_{{\alpha}}={\rm D}_{{\alpha}}K+{\bar{\rm D}}^2\Lambda_{{\alpha}} ~~~. \label{equ53}
\end{equation}
The change of the general action (\ref{equ52}) under this transformation is:
\begin{equation}
\eqalign{
\delta S=\int d^8z &\left(-2c_1{\rm D}_{{\alpha}}\Psi^{{\alpha}}+a_2{\bar{\rm D}}_{{\dot{\alpha}}}\bar{\Psi}^{{\dot{\alpha}}}\right){\rm D}^{{\beta}}
{\bar{\rm D}}^2\Lambda_{{\beta}}+c.c.\cr
&+2c_2\Psi^{{\alpha}}{\bar{\rm D}}^2{\rm D}_{{\alpha}}K+c.c.\cr
&-a_1\bar{\Psi}^{{\dot{\alpha}}}{\rm D}^2{\bar{\rm D}}_{{\dot{\alpha}}}K+c.c.\cr
&+\left(2a_2-a_1\right)\bar{\Psi}^{{\dot{\alpha}}}{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^2K+c.c.\cr
} \label{equ54}
\end{equation}
At this point, we can do a very useful observation. All the propagating degrees of freedom
required for the formulation of a massless $Y=1$ theory, can be included in the main
superfield $\Psi_{{\alpha}}$. To verify that, just look the Taylor expansion of the superfields
\eqref{taylorM} and \eqref{taylorA}. So as a consequence, either we have to make the
change of the action to vanish or add purely auxiliary compensators.
\subsection{A) ~~$K={\rm D}^{{\alpha}}U_{{\alpha}}$}
For this choice of $K$ we find the following action
\begin{equation}
\eqalign{
S=\int d^8z\Bigg\{
-&\frac{1}{2}a_1\Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} +c.c. \cr
+&a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar {\Psi}}_{{\dot{\alpha}}}\cr
+&a_1\left({\rm D}_{{\alpha}}{\bar{\rm D}}^2\Psi^{{\alpha}}+{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^2\bar{\Psi}^{{\dot{\alpha}}}\right)V\cr
+&\frac{1}{2}a_1V{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V\Bigg\}~~,
} \label{equ55}
\end{equation}
and this action is invariant under the transformations
\begin{equation}
\eqalign{
&\delta\Psi_{{\alpha}}=-{\rm D}^2U_{{\alpha}}+{\bar{\rm D}}^2\Lambda_{{\alpha}}~~,\cr
&\delta V={\rm D}^{{\alpha}}U_{{\alpha}}+{\bar{\rm D}}^{{\dot{\alpha}}}\bar{U}_{{\dot{\alpha}}}~~,
} \label{equ56}
\end{equation}
and the Bianchi Identities are:
\begin{equation}
\eqalign{
&{\rm D}^2{\bm{\cal T}}_{{\alpha}}+{\rm D}_{{\alpha}}{\bm{\cal P}}=0~~,\cr
&{\bar{\rm D}}^2{\bm{\cal T}}_{{\alpha}}=0~~.
} \label{equ57}
\end{equation}
where
\begin{equation}
\eqalign{
&{\bm{\cal T}}_{{\alpha}}=-a_1{\bar{\rm D}}^2\Psi_{{\alpha}}+a_1{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}\bar{\Psi}_{{\dot{\alpha}}}+a_1
{\bar{\rm D}}^2{\rm D}_{{\alpha}}V~~,\cr
&{\bm{\cal P}}=a_1{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V-a_1\left({\rm D}^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}}+{\bar{\rm D}}^{{\dot{\alpha}}}
{\rm D}^2\bar{\Psi}_{{\dot{\alpha}}}\right)~~.\cr
} \label{equ58}
\end{equation}
This is the $s=1$ limit of \eqref{equ19}.
On-shell the propagating degrees of freedom of superfield $V$ are gauged away completely
and the only thing that survives is the $Y=1$ supermultiplet. This can be visualized by the
following argument. There is a gauge where $V=0$. This happens for $U_{{\alpha}}=i{\bar{\rm D}}^2{\rm D}_{{\alpha}}
L$. Working in this gauge the $V$ superfield vanishes from the action which becomes:
\begin{equation}
\eqalign{
S=\int d^8z\Bigg\{
-&\frac{1}{2}a_1\Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} -\frac{1}{2}a_1\bar{\Psi}^{{\dot{\alpha}}}{\rm D}^2\bar{\Psi}_{{\dot{\alpha}}} +a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar {\Psi}}_{{\dot{\alpha}}}
\Bigg\}~~,
} \label{equ59}
\end{equation}
and is invariant under the transformation
\begin{equation}
\delta\Psi_{{\alpha}}={\rm D}^2{\bar{\rm D}}^{{\dot{\alpha}}}\pa_{{\alpha}{\dot{\alpha}}}L+{\bar{\rm D}}^2\Lambda_{{\alpha}}~~.
\label{equ60}
\end{equation}
This is the formulation suggested first by Fradkin and Vasiliev in \cite{Fradkin:1979as} and de
Wit and van Holten in \cite{deWvanH} at the component level and in \cite{Gates:1979gv} for a
superfield description. This last work also made the observation that this formulation was distinct
from an earlier off-shell description of the $Y$ $=$ 1 supermultiplet \cite{Ogievetsky:1975vk}.
\subsection{B) ~~$K={\bar{\rm D}}^{{\dot{\alpha}}}U_{{\dot{\alpha}}}$}
For this choice of $K$ we get the action:
\begin{equation}
\eqalign{
S=\int d^8z \Big\{
-&\frac{1}{2}a_1\Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} +c.c. \cr
+&a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar{\Psi}}_{{\dot{\alpha}}}\cr
+&a_1\left({\rm D}_{{\alpha}}{\bar{\rm D}}^2\Psi^{{\alpha}}+{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^2\bar{\Psi}^{{\dot{\alpha}}}\right)V\cr
+&a_1\left({\bar{\rm D}}^2{\rm D}_{{\alpha}}\Psi^{{\alpha}}+{\rm D}^2{\bar{\rm D}}_{{\dot{\alpha}}}\bar{\Psi}^{{\dot{\alpha}}}\right)V\cr
+&\frac{3}{2}a_1V{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V\cr
+&a_1V\Box V\Big\}~~,
} \label{equ61}
\end{equation}
and it is invariant under the gauge transformations
\begin{equation}
\eqalign{
&\delta\Psi_{{\alpha}}={\rm D}_{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}U_{{\dot{\alpha}}}+{\bar{\rm D}}^2\Lambda_{{\alpha}}~~,\cr
&\delta V={\bar{\rm D}}^{{\dot{\alpha}}}U_{{\dot{\alpha}}}+{\rm D}^{{\alpha}}\bar{U}_{{\alpha}} ~~,
} \label{equ62}
\end{equation}
with Bianchi Identities
\begin{equation}
\eqalign{
&{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^{{\alpha}}{\bm{\cal T}}_{{\alpha}}-{\bar{\rm D}}_{{\dot{\alpha}}}{\bm{\cal P}}=0\cr
&{\bar{\rm D}}^2{\bm{\cal T}}_{{\alpha}}=0
} \label{equ63}
\end{equation}
with
\begin{equation}
\eqalign{
&{\bm{\cal T}}_{{\alpha}}=-a_1{\bar{\rm D}}^2\Psi_{{\alpha}}+a_1{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}\bar{\Psi}_{{\dot{\alpha}}}+a_1\left\{{\bar{\rm D}}^2,{\rm D}_{{\alpha}}\right\}V~~,\cr
&{\bm{\cal P}}=3a_1{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V+2a_1\Box V-a_1\left[\left\{{\bar{\rm D}}^2,{\rm D}^{{\alpha}}\right\}\Psi_{{\alpha}}+\left\{{\rm D}^2,{\bar{\rm D}}^{{\dot{\alpha}}}\right\}\bar{\Psi}_{{\dot{\alpha}}}\right]~~.\cr
} \label{equ64}
\end{equation}
Like before there is no component of $V$ surviving on-shell.
The only propagating sub-multiplet is the $Y=1$ supermultiplet. This is the $s=1$
limit of \eqref{action}.
\subsection{C) One more thing...\\
${~~~~}K=\bar{K},~\Lambda_{{\alpha}}=i{\rm D}_{{\alpha}}U,~U=\bar{U}$}
For the special case of $s=1$ there is one more possibility.\\
If $K=\bar{K}$,~$\Lambda_{{\alpha}}=i{\rm D}_{{\alpha}}U$,~$U=\bar{U}$ the change of the
action becomes
\begin{equation}
\eqalign{
\delta S=\int d^8z &~ i\left(-2c_1-a_2\right){\rm D}_{{\alpha}}\Psi^{{\alpha}}{\rm D}^{{\beta}}{\bar{\rm D}}^2{\rm D}_{{\beta}}U+c.c.\cr
&+\left(2c_2-a_1\right)\Psi^{{\alpha}}{\bar{\rm D}}^2{\rm D}_{{\alpha}}K+c.c.\cr
&+\left(2a_2-a_1\right)\bar{\Psi}^{{\dot{\alpha}}}{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^2K+c.c.\cr
} \label{equ65}
\end{equation}
which suggests that by choosing:
\begin{equation}
\eqalign{
-2c_1&=a_2 ~~,~~
2c_2 =a_1~~,~~
2a_2 =a_1 ~~.
} \label{equ66}
\end{equation}
we get:
\begin{equation}
\eqalign{
S=\int d^8z \Big\{
&+a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar{\Psi}}_{{\dot{\alpha}}}
+\fracm{1}{2}a_1\Psi^{{\alpha}}{\rm D}_{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\Bar {\Psi}}_{{\dot{\alpha}}}\cr
&~+ a_1 \big[ ~ - \fracm 14 \Psi^{{\alpha}}{\rm D}^2\Psi_{{\alpha}}
~+ \fracm 12 \Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} +c.c. \big] ~
\Big\}
} \label{equ67}
\end{equation}
This is invariant under the transformation
\begin{equation}
\delta\Psi_{{\alpha}}={\rm D}_{{\alpha}}K+i{\bar{\rm D}}^2{\rm D}_{{\alpha}}U ~~, \label{equ68}
\end{equation}
with $K=\bar{K}$,~$U=\bar{U}$. This formulation was considered by Ogievetsky and
Sokatchev in \cite{Ogievetsky:1975vk} and was noted later in
\cite{Gates:1979gv} and describes a massless $Y=1$ supermultiplet. Referring back to this work, it can be seen that the following
spectrum of fields was presented.
\begin{center}
\begin{tabular} {| c | c | c |}
\hline
Component Field(s) & Bosonic & Fermionic\\
\hline
$A_{{\alpha} {\dot{\alpha}}}$ & $3$ & {}\\
\hline
$\psi_{{\alpha} {\beta} {\dot{\alpha}}}$~/~$\psi_{{\alpha} {\dot{\alpha}}}$ & {} & $ 12 $\\
\hline
$P$ & $1$ & {}\\
\hline
$\lambda_{{\dot{\alpha}}}$ & {} & $ 4 $\\
\hline
$Y_{{\alpha} {\dot{\alpha}} } $ & $ 8 $ & {}\\
\hline
$t_{{\alpha} {\beta} } $ & $6$ & {}\\
\hline
$M$ & $1$ & {}\\
\hline
$N$ & $1$ & {}\\
\hline
$\chi_{{\alpha}(s)}$ & {} & $4 $\\
\hline\hline
$~$ & $20$ & $20$\\
\hline
\end{tabular}
\end{center}
\noindent
A brief comparison between this table and the previous reveals a surprise, but a
very satisfying one.
To take the limit of the table at the bottom of page seventeen we begin
by substituting $s$ $=$ 1. Upon this substitution, any field with a subscript
that takes a 0-value means that index does not appear on the field. For any field
with a subscript that takes a value $<$ 0 means that field does not appear
at all. When these rules are applied and the value $s$ $=$ 1 is used in the
second and third columns, the two table match perfectly\footnote{We anticipated this
in naming the fields that appear in the expansions on page fifteen.}!
In other words, the matter gravitino multiplet described as ``the (3/2,1) superfield
of O(2) supergravity'' is the lowest member of the FVdWH-series tower of higher spin multiplets
of such theories. If this is true that means that there must be a duality between the two theories
in case B) and case C). The answer is yes, these theories are dual to each other and the
duality mechanism is provided by the $s=1$ limit of \eqref{equ35}
\begin{equation}
\eqalign{ {~~~~~}
S=\int d^8z \Big\{
-&\frac{1}{4}a_1\Psi^{{\alpha}}{\rm D}^2\Psi_{{\alpha}}-\frac{1}{2}a_1\Psi^{{\alpha}}{\bar{\rm D}}^2\Psi_{{\alpha}} +c.c. \cr
+&a_1\Psi^{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}_{{\alpha}}{\Bar{\Psi}}_{{\dot{\alpha}}}+\[\frac{1}{2}\]a_1\Psi^{{\alpha}}{\rm D}_{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}{\Bar {\Psi}}_{{\dot{\alpha}}}\cr
-&\[\frac{1}{2}\]a_1\left({\rm D}_{{\alpha}}\Psi^{{\alpha}}+{\bar{\rm D}}_{{\dot{\alpha}}}\bar{\Psi}^{{\dot{\alpha}}}\right)B\cr
+&a_1\left({\rm D}_{{\alpha}}{\bar{\rm D}}^2\Psi^{{\alpha}}+{\bar{\rm D}}_{{\dot{\alpha}}}{\rm D}^2\bar{\Psi}^{{\dot{\alpha}}}\right)V\cr
-&\frac{1}{4}a_1BB+\frac{1}{2}a_1V{\rm D}^{{\gamma}}{\bar{\rm D}}^2{\rm D}_{{\gamma}}V\cr
-&a_1B\left({\rm D}^2V+{\bar{\rm D}}^2V\right)\Big\} ~~~,
}
\end{equation}
which is invariant under the transformations
\begin{equation}
\eqalign{
&\delta\Psi_{{\alpha}}={\rm D}_{{\alpha}}{\bar{\rm D}}^{{\dot{\alpha}}}U_{{\dot{\alpha}}}-{\bar{\rm D}}^2\Lambda_{{\alpha}} ~~~,\cr
&\delta V={\bar{\rm D}}^{{\dot{\alpha}}}U_{{\dot{\alpha}}}+{\rm D}^{{\alpha}}\bar{U}_{{\alpha}} ~~,\cr
&\delta B=-{\rm D}^{{\alpha}}{\bar{\rm D}}^2\Lambda_{{\alpha}}-{\bar{\rm D}}^{{\dot{\alpha}}}{\rm D}^2\bar{\Lambda}_{{\dot{\alpha}}} ~~.
}
\end{equation}
From this point forward there are two choices:
\begin{itemize}
\item Choice 1) We can integrate out the auxiliary superfield B as we did in
the general case and this will give us \eqref{equ61}
\item Choice 2) We can work in a gauge where $B=V=0$ and this will give
(up to a redefinition of the gauge parameter) \eqref{equ67}
\end{itemize}\newpage
$$
\vCent
{\setlength{\unitlength}{1mm}
\begin{picture}(-20,-140)
\put(-55,-55){\includegraphics[width=4in]{Towers}}
\end{picture}}
$$
\vskip2.1in
In our accompanying work of the half-odd integer case \cite{Gates:2011qa},
we found that the non-minimal
off-shell supergravity theory first discovered by Breitenlohner \cite{nonmin1}, is the
lowest level of an infinite towers of such theories. In this work we have found
the same thing for ``the (3/2,1) superfield of O(2) supergravity.'' This particular
off-shell matter gravitino multiplet together with the non-minimal off-shell supergravity
multiplet provides a description of 4D, $\cal N$ $=$ 2 supergravity \cite{deWvanH,
N2SUSYn1SF}. It is therefore reasonable
to expect\footnote{A similar realization of 4D, $\cal N$ $=$ 2 supersymmetry has been found
previously \cite{N2}.} that this 4D, $\cal N$ $=$ 2 supersymmetry can
persist when the B-series (${\bm{\cal W}}_{{\alpha}(2s +1)}$, $ {\bm{{\cal G}}}_{{\alpha}(s){\dot{\alpha}}(
s)}$ and ${\bm{{\cal T}}}_{{\alpha}(s){\dot{\alpha}}(s-1)}$) and
FVdWH-series (${\bm{\cal W}}_{{\alpha}(2s)}$, $ {\bm{{\cal P}}}_{{\alpha}(s-1){\dot{\alpha}}(
s-1)}$ and ${\bm{{\cal T}}}_{{\alpha}(s){\dot{\alpha}}(s-1)}$) towers are taken together. A similar behavior
was observed for alternate towers \cite{N2}.
\vskip.2in
${~~~}$ \newline
${~~~~~}$``{\it {The opposite of a correct statement is a false statement. But the }}${~~~}$ \newline
${~~~~~~~}${\it {opposite
of a profound truth may well be another profound truth.
}}''
\newline $~~~~~~~$ -- Niels Bohr
\newline ${~~~}$
\noindent
{\Large\bf Acknowledgments}
This research was supported in part by the endowment of the John S.~Toll Professorship, the University of Maryland
Center for String \& Particle Theory, National Science Foundation Grant PHY-0354401. This work is also supported
by U.S. Department of Energy (D.O.E.) under cooperative agreement DEFG0205ER41360. SJG offers additional
gratitude to the M.\ L.\ K. Visiting Professorship program and to the M.\ I.\ T.\ Center for Theoretical Physics for support and
hospitality extended during the undertaking of this work.
\newpage
|
\section{Introduction}
Given any property ${\mathbf P}$ satisfied by the primes, it is natural to consider the set ${\mathcal C}_{\mathbf P}:=\{n~{\text{\rm composite}}: n~{\text{\rm satisfies}} ~{\mathbf P}\}$. Elements of ${\mathcal C}_{\mathbf P}$ can be thought of as pseudoprimes with respect to the property ${\mathbf P}$. Such sets of pseudoprimes have been of interest to number theorists.
Putting aside practical primality tests such as Fermat, Euler, Euler--Jacobi, Miller--Rabin, Solovay--Strassen, and others, let us have a look at some interesting, although not very efficient, primality tests as summarized in the table below.
\medskip
\begin{tabular}{|c|c|c|c|}
\hline
& Test & Pseudoprimes & Infinitely many \\ \hline
1 & $(n-1)!\equiv-1\pmod n$ & None & No \\ \hline
2 & $ a^n \equiv a\pmod n$ {\text{\rm for~all}} $a$ & Carmichael numbers &Yes \\ \hline
3 & $\sum_{j=1} ^ {n-1} j^{\phi(n)}\equiv -1 \pmod n$ & Giuga numbers & Unknown \\ \hline
4 & $\phi(n) | (n-1) $& Lehmer numbers & No example known \\ \hline
5 & $\sum_{j=1} ^ {n-1} j^{ n-1}\equiv -1 \pmod n $ & & No example known \\ \hline
\end{tabular}
\medskip
In the above table, $\phi(n)$ is the Euler function of $n$.
The first test in the table, due to Wilson and published by Waring in \cite{Wil}, is an interesting and impractical characterization of a prime number. As a consequence, no pseudoprimes for this test exist.
The pseudoprimes for the second test in the table are called Carmichael numbers. They were characterized by Korselt in \cite{Kor}. In \cite{AGP}, it is proved that there are infinitely many of them. The counting function for the Carmichael numbers was studied by Erd\H os in \cite{Erd} and by Harman in \cite{Har}.
The pseudoprimes for the third test are called Giuga numbers. The sequence of such numbers is sequence A007850 in OEIS. These numbers were introduced and characterized in \cite{Bor}. For example, a Giuga number is a squarefree composite integer $n$ such that $p$ divides $n/p-1$ for all prime factors $p$ of $n$. All known Giuga numbers are even. If an odd Giuga number exists, it must be the product of at least $14$ primes. The Giuga numbers also satisfy the congruence $nB_{\phi(n)}\equiv-1\pmod n$, where for a positive integer $m$ the notation $B_m$ stands for the $m$th Bernoulli number.
The fourth test in the table is due to Lehmer (see \cite{Leh}) and it dates back to 1932. Although it has recently drawn much attention, it is still not known whether any pseudoprimes at all exist for this test or not. In a series of papers (see \cite{Pom1}, \cite{Pom2}, and \cite{Pom3}), Pomerance has obtained upper bounds for the counting function of the Lehmer numbers, which are the pseudoprimes for this test. In his third paper \cite{Pom3}, he succeded in showing that the counting function of the Lehmer numbers $n\le x$ is $O(x^{1/2} (\log x)^{3/4})$. Refinements of the underlying method of \cite{Pom3} led to subsequent improvements in the exponent of the logarithm in the above bound by Shan \cite{Sha}, Banks and Luca \cite{BaL}, Banks, G\"ulo\u glu and Nevans \cite{BGN}, and Luca and Pomerance \cite{LP}, respectively. The best exponent to date is due to Luca and Pomerance \cite{LP} and it is $-1/2+\varepsilon$ for any $\varepsilon>0$.
The last test in the table is based on a conjecture formulated in 1959 by Giuga \cite{Giu}, which states that the set of pseudoprimes for this test is empty. In \cite{Bor}, it is shown that every counterexample to Giuga's conjecture is both a Carmichael number and a Giuga number. Luca, Pomerance and Shparlinski \cite{LPS} have showed that the counting function for these numbers $n\le x$ is $O(x^{1/2}/(\log x)^{2})$ improving slightly on a previous result by Tipu \cite{Tip}.
In this paper, inspired by Giuga's conjecture, we study the odd positive integers $n$ satisfying the congruence
\begin{equation}
\label{eq:newgiuga}
\sum_{j=1}^{n-1} j^{(n-1)/2}\equiv 0 \pmod n.
\end{equation}
It is easy to see that if $n$ is an odd prime, then $n$ satisfies the above congruence. We characterize such positive integers $n$ and show that they have an asymptotic density which turns out to be slightly larger than $3/8$.
For simplicity we put
$$
G(n)=\sum_{j=1}^{n-1} j^{\lfloor (n-1)/2\rfloor},
$$
although we study this function only for odd values of $n$.
\section{On the congruence $G(n)\equiv 0\pmod n$ for odd $n$}
We put
$$
{\mathfrak P}:=\{n~{\text{\rm odd}}: G(n)\equiv 0\pmod n\}.
$$
It is easy to observe that every odd prime lies in ${\mathfrak P}$. In fact, by Euler's criterion, if $p$ is an odd prime, then ${\displaystyle{j^{(p-1)/2}\equiv \left(\frac{j}{p}\right)\pmod p}}$, where ${\displaystyle{\left(\frac{j}{p}\right)}}$ denotes the Legendre symbol of $j$ with respect to $p$. Thus,
$$
G(p)\equiv \sum_{j=1}^{p-1} \left(\frac{j}{p}\right)\equiv 0\pmod p,
$$
so that $p\in {\mathfrak P}$.
We start by showing that numbers which are congruent to $3\pmod 4$ are in ${\mathfrak P}$.
\begin{prop} If $n\equiv 3\pmod 4$, then $n \in {\mathfrak P} $.
\end{prop}
\begin{proof}
Writing $n=4m+3$, we have that $(n-1)/2=2m+1$ is odd. Now,
\begin{eqnarray*}
2G(n) & = & \sum_{j=1}^{n-1} \left(j^{2m+1}+(n-j)^{2m+1}\right)\\
& = & n\sum_{j=1}^{n-1} \left(j^{2m}+j^{2m-1}(n-j)+\cdots+(n-j)^{2m}\right),
\end{eqnarray*}
so $n\mid 2G(n)$. Since $n$ is odd, we get that $G(n)\equiv 0\pmod n$, which is what we wanted.
\end{proof}
The next lemma is immediate.
\begin{lemma}
\label{lem:1}
Let $p$ be an odd prime and let $k\geq1$ be an integer. Then
$$\gcd\left(\frac{p^k-1}{2},\varphi(p^k)\right)=\gcd\left(\frac{p^k-1}{2},p-1\right)=\begin{cases} p-1 & \textrm{if $k$ is even,}\\
(p-1)/2 & \textrm{if $k$ is odd.}
\end{cases}$$
\end{lemma}
With this lemma in mind we can prove the following result.
\begin{prop}
\label{prop:1}
Let $p$ be an odd prime and let $k\geq 1$ be any integer. Then, $p^k\in{\mathfrak P}$ if and only if $k$ is odd.
\end{prop}
\begin{proof}
Let $\alpha\in \mathbb{Z}$ be an integer whose class modulo $p^k$ is a generator of the unit group of $\mathbb{Z}/p^k\mathbb{Z}$. We put $\beta:=\alpha^{(p^k-1)/2}$. Suppose first that $k$ is odd. We then claim that $\beta-1$ is not zero modulo $p$. In fact, if $\alpha^{(p^k-1)/2}\equiv 1\pmod p$, then since also $\alpha^{p-1}\equiv 1\pmod p$, we get, by Lemma \ref{lem:1}, that $\alpha^{(p-1)/2}\equiv 1\pmod p$, which is impossible.
Now, since $\beta-1$ is coprime to $p$, it is invertible modulo $p^k$. Moreover, since also $k\leq (p^k-1)/2$, we have that
\begin{eqnarray*}
G(n) & = & \sum_{j=1}^{n-1} j^{(p^k-1)/2}\equiv \sum_{\substack{\gcd(j,p)=1\\ 1\leq j\leq n-1}} j^{(p^k-1)/2}\pmod {p^k}\\
& \equiv & \sum_{j=1}^{\varphi(p^k)}\left(\alpha^{(p^k-1)/2}\right)^i\pmod {p^k}\equiv \sum_{i=1}^{\phi(p^k)} \beta^{i} \pmod {p^k}\\
& = & \frac{\beta^{\varphi(p^k)+1}-\beta}{\beta-1}\equiv 0\pmod {p^k}.
\end{eqnarray*}
Assume now that $k$ is even. Observe that
$$
(p^k-1)/2=(p-1)((1+p+\cdots+p^{k-1})/2):=(p-1)m,
$$
and $m$ is an integer which is coprime to $p$. Thus, $\beta=\alpha^{(p^k-1)/2}=(\alpha^{(p-1)})^m$ has order $p^{k-1}$ modulo $p^k$, and so does
$\alpha^{p-1}$. Moreover, again since $k\le (p^k-1)/2$, we may eliminate the multiples of $p$ from the sum defining $G(n)$ modulo $n$ and get
\begin{eqnarray}
\label{eq:kiseven}
G(n) & = & \sum_{j=1}^{n-1} j^{(p^k-1)/2}\equiv \sum_{\substack{\gcd(j,p)=1\\ 1\le j\le n-1}} j^{(p^k-1)/2}\pmod {p^k}\nonumber\\
& \equiv & \sum_{i=1}^{\varphi(p^k)}\left(\alpha^{(p^k-1)/2}\right)^i\equiv \sum_{i=1}^{p^{k-1}(p-1)}\left(\alpha^{(p-1)}\right)^{im}\pmod {p^k} \nonumber\\
& \equiv & (p-1)\sum_{i=1}^{p^{k-1}}\left(\alpha^{p-1}\right)^i\pmod {p^k}.
\end{eqnarray}
Since $\alpha^{p-1}$ has order $p^{k-1}$ modulo $p^k$, it follows that $\alpha^{p-1}=1+pu$ for some integer $u$ which is coprime to $p$. Then
\begin{equation}
\label{eq:2}
\sum_{i=1}^{p^{k-1}}\left(\alpha^{p-1}\right)^i=\alpha \left(\frac{\alpha^{p^{k-1}}-1}{\alpha-1}\right).
\end{equation}
Since $\alpha^{p^{k-1}}\equiv 1+p^k u\pmod {p^{k+1}}$, it follows that $(\alpha^{p^{k-1}}-1)/(\alpha-1)\equiv p^{k-1}\pmod {p^k}$, so that
\begin{equation}
\label{eq:3}
\alpha\left(\frac{\alpha^{p^{k-1}}-1}{\alpha-1}\right)\equiv \alpha p^{k-1}\pmod {p^k}\equiv p^{k-1}\pmod {p^k}.
\end{equation}
Calculations \eqref{eq:2} and \eqref{eq:3} together with congruences \eqref{eq:kiseven} give that $G(n)\equiv (p-1)p^{k-1}\pmod {p^k}$. Thus,
$p^k$ is not in ${\mathfrak P}$ when $k$ is even.
\end{proof}
Note that Proposition \ref{prop:1} does not extend to powers of positive integers having at least two distinct prime factors. For example,
$n=2021=43\times 47$ has the property that both $n$ and $n^2$ belong $\mathfrak{P}$.
\section{A characterization of $\mathfrak{P}$ and applications}
Here, we take a look into the arithmetic structure of the elements lying in $\mathfrak{P}$. We start with an easy but useful lemma.
\begin{lemma}
\label{lem:2}
Let $n=\prod_{p^{r_p}\| n} p^{r_p}$ be an odd integer, and let $A$ be any positive integer. If $\gcd(A,p-1)<p-1$ for all $p\mid n$, then
$$
\sum_{\substack{\gcd(j,n)=1\\1\leq j\leq n-1}}j^A\equiv 0\pmod n.
$$
\end{lemma}
\begin{proof}
It suffices to prove that the above congruence holds for all prime powers $p^{r_p}\| n$. So, let $p^{r}$ be such a prime power and let $\alpha$ be an
integer which is a generator of the unit group of $\mathbb{Z}/p^{r}\mathbb{Z}$.
Put $\beta:=\alpha^A$. An argument similar to the one used in the proof of Proposition \ref{prop:1} (the case when $k$ is odd) shows that the condition
$\gcd(A,p-1)<p-1$ entails that $\beta-1$ is not a multiple of $p$. Thus, $\beta-1$ is invertible modulo $p$. We now have
\begin{eqnarray*}
\sum_{\substack{\gcd(j,n)=1\\ 1\le j\le n-1}} j^{A} & \equiv & \left(\frac{\phi(n)}{\phi(p^{r})}\right)\sum_{\substack{\gcd(j,p)=1\\ 1\le j\le p}} j^{A}\pmod {p^r}
\equiv \phi(n/p^r)\sum_{i=1}^{\phi(p^r)} \alpha^{Ai}\pmod {p^r}\\
& \equiv & \phi(n/p^r)\sum_{i=1}^{\phi(p^r)} \beta^i\pmod {p^r} \equiv \phi(n/p^r)\frac{\beta^{\phi(p^r)+1}-\beta}{\beta-1}\pmod {p^r}\\
& \equiv & 0\pmod {p^r},
\end{eqnarray*}
which is what we wanted to prove.
\end{proof}
\begin{theorem}
\label{thm:1}
A positive integer $n$ is in ${\mathfrak P}$ if and only if $\gcd((n-1)/2,p-1)<p-1$ for all $p\mid n$.
\end{theorem}
\begin{proof}
Assume that $n$ is odd and $\gcd((n-1)/2,p-1)<p-1$. By Lemma \ref{lem:2},
$$
\sum_{\substack{(j,n)=1\\ 1\leq j\leq n-1}} j^{(n-1)/2}\equiv 0 \pmod n.
$$
Now, let $d$ be any divisor of $n$. Observe that
\begin{equation}
\label{eq:d}
\sum_{\substack{(j,n)=d\\ 1\leq j\leq n-1}}j^{\frac{n-1}{2}}=d^{\frac{n-1}{2}}\sum_{\substack{(i,n/d)=1\\ 1\leq i\leq n/d-1}}i^{\frac{n-1}{2}}.
\end{equation}
The last sum in the right--hand side of \eqref{eq:d} above is, by Lemma \ref{lem:2}, a multiple of $n/d$, so that the sum in the left--hand side of
\eqref{eq:d} above
is a multiple of $n$. Summing up these congruences over all possible divisors $d$ of $n$ and noting that
$$
G(n)=\sum_{d\mid n} \sum_{\substack{\gcd(j,n)=d\\ 1\le j\le n-1}} j^{(n-1)/2},
$$
we get that $G(n)\equiv 0\pmod n$, so $n\in {\mathfrak P}$.
Conversely, say $n\in {\mathfrak P}$ is some odd number and assume that there exists a prime factor $p$ of $n$ such that $p-1\mid (n-1)/2$. Write $(n-1)/2=(p-1)m$. Observe that $m$ is coprime to $p$. Assume that $p^{r}\|n$. Then, modulo $p^r$, we have
$$
G(n) =\sum_{j=1}^{n-1} j^{(n-1)/2}\equiv (n/p^r) \sum_{\substack{\gcd(j,p)=1\\ 1\le j\le p^r-1}} j^{(n-1)/2}\pmod {p^r}\equiv
(n/p^r)\sum_{\substack{\gcd(j,p)=1\\ 1\le j\le p^r-1}} j^{(p-1)}.
$$
The argument used in Proposition \ref{prop:1} (the case when $k$ is even), shows that the second sum is not zero modulo $p^r$, and since $n/p^r$ is also coprime to $p$, we get that $p^r$ does not divide $G(n)$, a contradiction.
This completes the proof of the theorem.
\end{proof}
Here are a few immediate corollaries of Theorem \ref{thm:1}.
\begin{cor}
Let $n$ be any integer. Assume that one of the following conditions hold:
\begin{itemize}
\item[i)] $\gcd\left((n-1)/2,\varphi(n)\right)$ is odd;
\item[ii)] $\gcd\left((n-1)/2,\lambda(n)\right)$ is odd, where $\lambda(n)$ the Carmichael function.
\end{itemize}
Then $n\in {\mathfrak P}$.
\end{cor}
\begin{cor}
If $n^k\in {\mathfrak P}$ for some $k\geq 1$, then $n\in {\mathfrak P}$.
\end{cor}
\begin{proof}
Observe that $\gcd\left((n-1)/2,p-1\right)$ divides $\gcd\left((n^k-1)/2,p-1\right)$ for every $k$ and every prime number $p$.
Now the corollary follows from Theorem \ref{thm:1}.
\end{proof}
We add another sufficient condition which is somewhat reminiscent of the characterization of the Giuga numbers.
\begin{prop}
\label{prop:2}
Let $n=\prod_{p^{r_p}\| n} p^{r_p}$ be an odd integer. If $p-1$ does not divide $n/p^{r_p}-1$ for every prime factor $p$ of $n$, then $n\in {\mathfrak P}$.
\end{prop}
\begin{proof}
By Theorem \ref{thm:1}, if $n\not\in {\mathfrak P}$, then there exists a prime factor $p$ of $n$ such that $p-1$ divides $(n-1)/2$. In particular, $p-1\mid n-1$. Since $p-1$ also divides $p^{r_p}-1$, it follows that $p-1$ divides $n-p^{r_p}=p^{r_p}(n/p^{r_p}-1)$. Since $p-1$ is obviously coprime to $p^{r_p}$, we get that $p-1$ divides
$n/p^{r_p}-1$, which is a contradiction.
\end{proof}
It is also easy to determine whether numbers of the form $2^m+1$ are in ${\mathfrak P}$. Indeed, assume that $2^m+1\not\in {\mathfrak P}$ for some
positive integer $m$. Then, by Theorem
\ref{thm:1}, there is some prime $p\mid 2^m+1$ such that $p-1\mid ((2^m+1)-1)/2=2^{m-1}$. Thus, $p=2^a+1$ for some $a\le m-1$, and so $p$ is a Fermat prime. In particular, $a=2^{\alpha}$ for some $\alpha\ge 0$. Since $p=2^{2^{\alpha}}+1$ is a proper divisor of $2^m+1$, it follows that $2^{\alpha}\mid m$ and $m/2^{\alpha}$ is odd. This is possible only when $2^{\alpha}$ is the exact power of $2$ in $m$ and $m$ is not a power of $2$. So, we have the following result.
\begin{prop}
Let $n=2^m+1$ and $m=2^{\alpha} m_1$ with $\alpha\ge 0$ and odd $m_1>1$. Then $n\in {\mathfrak P}$ unless $2^{2^{\alpha}}+1$ is a Fermat prime.
\end{prop}
\section{Asymptotic density of $\mathfrak{P}$}
Let ${\mathbb{I}}$ be the set of odd positive integers. In order to compute the asymptotic density of ${\mathfrak P}$, or to even prove that it exists, it suffices to understand the elements in its complement $\mathbb{I}\backslash {\mathfrak P}$. It turns out that this is easy. For an odd prime $p$ let
$$
{\mathcal F}_p:=\{p^2 \pmod {2p(p-1)}\}.
$$
Observe that ${\mathcal F}_p\subseteq {\mathbb{I}}$.
\begin{theorem}
\label{thm:3}
We have
\begin{equation}
\label{eq:complement}
\mathbb{I}\backslash {\mathfrak P}=\bigcup_{p\ge 3} {\mathcal F}_p.
\end{equation}
\end{theorem}
\begin{proof}
By Theorem \ref{thm:1}, we have that $n\not\in {\mathfrak P}$ if and only if $p-1$ divides $(n-1)/2$ for some prime factor $p$ of $n$. This condition is equivalent to $n\equiv 1\pmod {2(p-1)}$. Write $n=pm$ for some positive integer $m$. Since $p$ is invertible modulo $2(p-1)$, it follows that $m$ is uniquely determined modulo $2(p-1)$. It suffices to notice that the class of $m$ modulo $2(p-1)$ is in fact $p$ since then $pm\equiv p^2\equiv 1\pmod {2(p-1)}$ with the last congruence following because
$p^2-1=(p-1)(p+1)$ is a multiple of $2(p-1)$. This completes the proof.
\end{proof}
Observe that ${\mathcal F}_p$ is an arithmetic progression of difference $1/(2p(p-1))$. Since the series
$$
\sum_{p\ge 3} \frac{1}{2p(p-1)}
$$
is convergent, it follows immediately that $\mathbb{I}\backslash {\mathfrak P}$; hence, also ${\mathfrak P}$, has a density. This also suggests a way to compute the density of ${\mathfrak P}$ with arbitrary precision. Namely, say $\varepsilon>0$ is given. Let $3=p_1<p_2<\cdots$ be the increasing sequence of all the odd primes. Let $k:=k(\varepsilon)$ be minimal such that
$$
\sum_{j\ge k} \frac{1}{2p_j(p_j-1)}<\varepsilon.
$$
It then follows that numbers $n\not\in {\mathfrak P}$ which are divisible by a prime $p_j$ with $j\ge k$ belong to $\bigcup_{j\ge k} {\mathcal F}_{p_j}$, which is a set of density $<\varepsilon$. Thus, with an error of at most $\varepsilon$, the density of the set $\mathbb{I}\backslash {\mathfrak P}$ is the same as the density of
$$
\bigcup_{j<k} {\mathcal F}_{p_j},
$$
which is, by the Principle of Inclusion and Exclusion,
\begin{equation}
\label{eq:s}
\sum_{s\ge 1} \sum_{1\le i_1<i_2<\cdots<i_s\le k-1}\frac{\varepsilon_{i_1,i_2,\ldots,i_s}}{\lcm[2p_{i_1}(p_{i_1}-1),\ldots, 2p_{i_s}(p_{i_s}-1)]},
\end{equation}
with the coefficient $\varepsilon_{i_1,i_2,\ldots,i_s}$ being zero if $\bigcap_{t=1}^s {\mathcal F}_{p_{i_t}}=\emptyset$, and being $(-1)^{s-1}$ otherwise.
Taking $ \varepsilon:=0.00082$, we get that $k = 29$,
$$
\rho(\bigcup_{j<29} {\mathcal F}_{p_j})=\frac{274510632303283394907222287246970994037}{2284268907516688397400621108446881752020}\approx 0.120174,
$$
and consequently $\rho(\mathfrak{P})$ belongs to [0.379005, 0.379826]. So, we can say that
$$
\rho(\mathfrak{P}) = 0.379\ldots
$$
Here and in what follows, for a subset ${\mathcal A}$ of the set of positive integers we used $\rho({\mathcal A})$ for its density when it exists.
These computations were carried out with \emph{Mathematica}, for which it was necessary to have a good criterion to determine when the intersection of $\mathcal{ F}_p$ for various odd primes $p$ is empty. We devote a few words on this issue. Let us observe first that the condition $n\in {\mathcal F}_p$, which is equivalent
to the fact that $p\mid n$ and $p-1$ divides $(n-1)/2$, can be formulated as the pair congruences
\begin{eqnarray}
\label{eq:pair}
n & \equiv & 1\pmod {2(p-1)};\nonumber\\
n & \equiv & 0 \pmod p.
\end{eqnarray}
Assume now that ${\mathcal P}$ is some finite set of primes. Let us look at $\bigcap_{p\in {\mathcal P}} {\mathcal F}_p$. Put $m:=\prod_{p\in {\mathcal P}} p$. The first set of congruences \eqref{eq:pair} for all $p\in {\mathcal P}$ is equivalent to
\begin{equation}
\label{eq:cong1}
n\equiv 1 \pmod {2\lambda(m)},
\end{equation}
where $\lambda(m)=\lcm[p-1:p\in {\mathcal P}]$ is the Carmichael $\lambda$-function of $m$. The second set of congruences for $p\in {\mathcal P}$ is equivalent to
\begin{equation}
\label{eq:cong2}
n\equiv 0\pmod m.
\end{equation}
Since $1$ is not congruent to $0$ modulo any prime $q$, it follows that a necessary condition for \eqref{eq:cong1} and \eqref{eq:cong2} to hold simultaneously is that $m$ and $2\lambda(m)$ are coprime. This is also sufficient by the Chinese Remainder Lemma in order for the pair of congruences \eqref{eq:cong1} and \eqref{eq:cong2} to have a solution $n$. Since $m$ is also squarefree, the condition that $m>1$ is odd and $m$ and $2\lambda(m)$ are coprime is equivalent to $m>2$ and $m$ and $\phi(m)$ are coprime. Put
\begin{equation}
\label{eq:M}
{\mathcal M}:=\{m>2: \gcd(m,\phi(m))=1\}.
\end{equation}
Thus, we proved the following result.
\begin{prop}
Let ${\mathcal P}$ be a finite set of primes and put $m:=\prod_{p\in {\mathcal P}} p$. Then $\bigcap_{p\in {\mathcal P}} {\mathcal F}_p$ is nonempty if and only if
$m\in {\mathcal M}$, where this set is defined at \eqref{eq:M} above. If this is the case, then the set $\bigcap_{p\in {\mathcal P}} {\mathcal F}_p$ is an arithmetic progression of difference $1/(2m\lambda(m))$.
\end{prop}
The condition that $m\in {\mathcal M}$ can also be formulated by saying that $m$ is odd, squarefree and $p\nmid q-1$ for all primes $p$ and $q$ dividing $m$.
We recall that the set ${\mathcal M}$ has been studied intensively in the literature. For example, putting ${\mathcal M}(x)={\mathcal M}\cap [1,x]$, Erd\H os \cite{Erd1} proved that
$$
\#{\mathcal M}(x)=e^{-\gamma} (1+o(1))\frac{x}{\log\log\log x}\qquad {\text{\rm as}}\quad x\to \infty.
$$
In particular, it follows that if ${\mathcal P}$ is a finite set of primes, then $\bigcap_{p\in {\mathcal P}} {\mathcal F}_p\ne \emptyset$ if and only if ${\mathcal F}_p\bigcap {\mathcal F}_q\ne \emptyset$ for any two elements $p$ and $q$ of ${\mathcal P}$.
Finally, let us observe that with this formalism and the Principle of Inclusion and Exclusion, as in \eqref{eq:s} for example, we can write that
$$
\rho({\mathfrak P})=\sum_{m\in {\mathcal M}\cup \{1\}} \frac{(-1)^{\omega(m)}}{2m\lambda(m)}.
$$
Here, $\omega(m)$ is the number of distinct prime factors of $m$. The fact that the above series converges absolutely follows easily from the inequality $\lambda(m)>(\log m)^{c\log\log\log m}$ which holds with some positive constant $c$ for all sufficiently large $m$ (see \cite{EPS}), as well the fact that the series
$$
\sum_{m\ge 2} \frac{1}{m(\log m)^2}
$$
converges. We give no further details.
|
\section{General introductory remarks}
Authors are invited to observe the suggestions described in the
present document in order to ensure consistency and uniformity in the
style and layout of CERN PH preprints.
Although the typesetting rules that are described hare are of a
general nature, the present document explains their use with the
\texttt{cernphprep} \LaTeX{} class. This class file becomes
automatically available when running with the latest \LaTeX{} on
CERN's central \texttt{lxplus} system\footnote
{See \url{http://cern.ch/XML/textproc/texlivelinux2009.html} for
an explanation on modifying your \texttt{PATH} environment variable
correctly.}.
As the text frame is defined as $16\Ucm \times 24\Ucm$ (\ie a A4 sized
paper has $3\Ucm$ margins at the top and at the bottom, and $3\Ucm$
at the inner and $2\Ucm$ at the outer edge), please take care to
remain inside these limits. This is especially true for figures and
tabular content.
Note that it is the responsibility of the authors to obtain permission
from the copyright holder if material taken from other sources is
included in the source submitted as a CERN PH preprint.
\section{The title page}
Your CERN PH preprint \textbf{must always} begin with a title page. The
information contained on the title page is specified inside a
\texttt{titlepage} environment.
The minimal mandatory information that a title page must contain is
the following (as, \eg for the present document).
\begin{Verbatim}[numbers=left,fontsize=\small]
\documentclass{cernphprep}
\usepackage{cite,fancyvrb}
\begin{document}
\begin{titlepage}
\PHnumber{2010--xxx}
\PHdate{DD Month 2010}
\EXPnumber{IT-UDS-HUS-001}
\EXPdate{23 June 2010}
\DEFCOL{CDS-Library}
\title{Preparing a CERN PH preprint}
\author{Michel Goossens/IT-UDS\thanks{\texttt{Email:<EMAIL>}}}
\begin{abstract}
This document contains instructions for authors who want to prepare
a CERN PH Preprint using the \texttt{cernphprep} \LaTeX{} class.
...
\end{abstract}
\end{titlepage}
\end{Verbatim}
On line 2 you see how you can specify supplementary packages to use in
your document (here, the \texttt{cite} and \texttt{fancyvrb} packages
are loaded). The commands \verb!\PHnumber! and \verb!\PHdate! (lines
5 and 6) specify the number and the date that have been assigned by
the PH Departmental Office to your CERN PH preprint. You can add a
number and date for your experiment by using the \verb!\EXPnumber! and
\verb!\EXPdate! commands (lines 7 and 8). By default these are
empty. Then follow (lines 9 and 10) the title and the author(s)
(inside, resp., \verb!\title! and \verb!\author! commands). Finally,
the \texttt{abstract} environment (lines 11--15) lets you give a short
description of the content of the preprint. The command
\verb!\CERNCopyright! contains the copyright message that can appear
at the bottom of the front page (default: no message).
\subsection{Controling the running titles}
For the body of the text the running titles will show the (beginning
of the) title on the odd (right-hand) pages and the authors on the
even (left-hand) pages.
If the way the authors or the title which are printed in the running
titles is not what you want, you can control the typeset text by
specifying what you want inside the \texttt{titlepage} environment as
argument to a \verb!\ShortAuthor! or \verb!\Shorttitle! command,
respectively.
\subsection{Special setup for the LHC experiments}
For the LHC experiments, the experiment logo can also be
printed. Experiments can contact me to have their logo registered so
that their experiment's name is recognized as an option by the class
file.
\begin{Verbatim}[numbers=left,fontsize=\small]
\documentclass[CMS]{cernphprep}
\begin{document}%
\begin{titlepage}
\EXPnumber{QCD-09-010}
\PHnumber{...}
\PHdate{...}
\title{Full title for paper...}
\Collaboration{The CMS Collaboration%
\thanks{See Appendix~\ref{app:collab} for the list of collaboration
members}}
\ShortAuthor{The CMS Collaboration}
\ShortTitle{Short title}
\begin{abstract}
...
\end{abstract}
\end{titlepage}
\end{Verbatim}
The \texttt{[CMS]} option specified on the \verb!\documentclass!
command (line~1) signals that we are dealing with a preprint prepared
by the CMS experiment. Thus the CMS logo will also appear on the title
page. We note the presence of the \verb!\Collaboration! command,
which substitutes for the \verb!\author! command (lines~8--10) with a
reference to the appendix (via the \verb!\thanks! command) which
contains the names of the many authors of that paper. Note also the
\verb!\ShortAuthor! and \verb!\ShortTitle! commands define the text
for the running headers.
\subsection{Alternative ways of specifying author lists}
For experiments with a smaller number of authors, the
\texttt{cernphprep} class offers two ways of specifying the authors
and affiliations on the title page, as shown in Fig.~\ref{fig:title}.
The \LaTeX{} sources for the layouts shown at the left and right in
Fig.~\ref{fig:title} correspond to the example files
\texttt{cernphprepexa.tex} and \texttt{cernphprepexax.tex}, respectively.
\begin{figure}
\centering
\fbox{\includegraphics[width=.46\linewidth]{cernphprepexa1.pdf}}
\hfill
\fbox{\includegraphics[width=.46\linewidth]{cernphprepexax1.pdf}}
\caption{Two possible layouts of authors and affiliations on the title page}
\label{fig:title}
\end{figure}
\section{The document body}
\subsection{Sectioning commands and paragraphs}
\label{sec:sections}
The standard \LaTeX{} commands \verb!\section! (level 1),
\verb!\subsection! (level 2), \verb!\subsubsection! (level 3),
\verb!\paragraph! (level 4), should be used for headings.
To start a new paragraph it is sufficient to insert a blank line,
hence it is unnecessary to include \verb!\par! commands in your
\LaTeX{} source.
Please minimize explicit line break (\verb!\newline!, \verb!\\!) or
page break (\verb!\newpage!) commands and use them only when
\emph{absolutely} necessary and when the document is in its final
state.
\subsection{Equations}
\label{sec:equations}
An unnumbered single formula is delimited by \verb!\[! and \verb!\]!
(alternatively a \texttt{displaymath} environment can be used), while a
numbered equation is generated with the \texttt{equation} environment.
You should \textbf{never} use the \texttt{\$\$} construct.
For cross-referencing equations use the \LaTeX{} commands \verb!\label!
and \verb!\ref!, as explained in Section~\ref{sec:crossref}.
A simple example, with a cross-reference to a formula, follows.
\begin{Verbatim}[numbers=left,fontsize=\small]
Einstein has expressed the relation between energy $E$
and mass $m$ in his famous formula~(\ref{eq:einstein}):
\begin{equation}
E=mc^2~.
\label{eq:einstein}
\end{equation}
\end{Verbatim}
When typeset, this gives:
Einstein has expressed the relation between energy $E$ and mass $m$ in
his famous formula~(\ref{eq:einstein}):
\begin{equation}
E=mc^2~.
\label{eq:einstein}
\end{equation}
Equations should be treated as part of the text, and therefore
punctuated (with a space, \eg \verb!~!, between the end of the
equation and the punctuation mark, as shown in
Eq.~(\ref{eq:einstein}). Equations are numbered consecutively
throughout the report.
The CERN \LaTeX{} classes automatically load the \texttt{amsmath}
class, which offers a large choice of constructs for typesetting
mathematics~\cite{bib:voss2005}. See also Appendix~\ref{app:amsmath}
for some hints about using the \texttt{amsmath} extensions in an
optimal way.
\subsection{Figures}
\label{sec:figures}
Figures should be prepared electronically. The image should be of high
quality when printed (in black and white). In particular, all lines
should be heavy enough to be visible and text large enough to be
legible when the figure is printed \emph{at the final size}. If using
colour, the figure should print clearly in greyscale\footnote
{Beware of faint colours, such as yellow or light blue or red, which
are barely visible when reproduced as level of grey.}.
When running \texttt{latex} EPS (Encapsulated PostScript) files are to
be provided, while when running \texttt{pdflatex} PDF, PNG or JPEG
formats can be used.
As already mentioned in the introduction, before using material such
as illustrations taken from other sources, you must obtain permission
from the copyright holder.
All figures must remain within the page area ($16\Ucm \times 24\Ucm$),
where, if necessary, the page may be turned $90^\circ$ to accommodate
the figure. When this is done, the caption must be oriented in the
same way as the figure, and no other text may appear on that page.
The bottom of the turned illustrations should be at the right-hand
side of the page.
\subsubsection{Including your figure}
Figures are included with the \texttt{figure} environment. In the
example that follows we include an EPS graphics file
(\texttt{myfig.eps}, where the extension \texttt{.eps} does not need
to be specified since it is the default) with the
\verb!\includegraphics! command, which is defined in the
\texttt{graphicx} package that is loaded by default by the CERN
\LaTeX{} classes.
The figure caption (specified as argument of the \verb!\caption!
command), must \emph{follow} the figure body, and should be brief. No
full stop is necessary unless the caption is more that one sentence
long, in which case full punctuation should be used.
\subsubsection{References to figures}
\LaTeX's cross-reference mechanism can be used to refer to figures. A
figure reference is defined with a \verb!\label! command, which must
come in the source \emph{after} the \verb!\caption! command that
contains the description of the figure content.
Figures must be referenced in the text in consecutive numerical order
with the help of the \verb!\ref! command. The following are examples
of references to figures and how to produce them.
\begin{Itemize}
\item `Fig. 3' produced by, \eg \verb!Fig.~\ref{fig:myfig}!,
\item `Figs. 3--5' produced by, \eg
\verb!Figs.~\ref{fig:myfiga}--\ref{fig:myfigb}!,
\item At the beginning of a sentence the word ``Figure'' should be
written out in full. For instance, \verb!Figure~\ref{fig:myfig}!
which gives, \eg `Figure~3'.
\end{Itemize}
Figures with several parts are cited as follows: `Figs. 2(a) and (b)',
and `Figs. 3(a)--(c)'.
Figures and illustrations \emph{should follow} the paragraph in which
they are first discussed. If this is not feasible, they may be placed
on the following page (\LaTeX{}'s float mechanism takes care of this
automatically, in principle). If it is not possible to place
\emph{all} numbered figures in the text, then they should \emph{all}
be placed at the end of the paper.
An example with a cross-reference to a figure follows. The reference
is defined by the \verb!\label! command \emph{following} the
\verb!\caption! command. A cross-reference is generated with the
\verb!\ref! command on the second line. Note the cross-reference key
(\texttt{fig:myfig}) which clearly indicates that it refers to a
figure (see Section~\ref{sec:crossref} for a discussion of the importance of
using good keys).
\begin{Verbatim}[numbers=left,fontsize=\small]
\section{Section title}
In Fig.~\ref{fig:myfig} we see that ...
\begin{figure}
\centering\includegraphics[width=.9\linewidth]{myfig}
\caption{Description of my figure}
\label{fig:myfig}
\end{figure}
\end{Verbatim}
\subsection{Tables}
Tables are defined the the \texttt{table} environment. Each table
should be centred on the page width, with a brief caption (specified
as argument of the \verb!\caption! command) \emph{preceding} the table
body.
In general, tables should be open, drawn with a double thin horizontal
line (0.4~pt) at the top and bottom, and single horizontal line
(0.4~pt) separating column headings from data.
Like figures, tables must be referenced in the text in consecutive
numerical order with \LaTeX's \verb!\ref! command. Examples of
cross-references to tables follow (commands that can be used to
generate the given text strings are shown between parentheses): `Table
5' (\eg \verb!\Table~\ref{tab:mytab}!),
`Tables 2--3' (\eg
\verb!\Tables~\ref{tab:mytaba}--\ref{tab:mytabb}!). The word `Table'
should never be abbreviated.
\subsubsection{Formatting and layout within the table}
Write the headings in sentence case but do not use full stops. Units
should be entered in parentheses on a separate line below the column
heading. (If the same unit is used throughout the table, it should be
written in parentheses on a separate line below the caption.)
Unsimilar items should be aligned on the left, whereas similar items
should be aligned on the operator or decimal point. All decimal points
must be preceded by a digit.
\emph{Table captions} should be brief and placed centrally
\emph{above} the table. No full stop is necessary unless the caption
is more that one sentence long, in which case full punctuation should
be used.
\emph{Notes in tables} should be designated by superscript lower-case
letters, and begun anew for each table. The superscript letter should
be placed in alphabetical order moving from left to right across the
first row and down to the last. The notes should then be listed
directly under the table.
An instance of a simple table, following the proposed rules follows.
\begin{Verbatim}[numbers=left,fontsize=\small]
Table~\ref{tab:famous} shows three famous mathematical constants.
\begin{table}
\caption{Famous constants}
\label{tab:famous}
\centering
\begin{tabular}{@{}lll@{}}\hline\hline
Symbol & Description & Approximate value\\\hline
$e$ & base of natural logarithm & $2.7182818285$ \\
$\pi$ & ratio circle circumference to diameter &
$3.1415926536$ \\
$\phi$ & golden ratio & $1.6180339887$ \\\hline\hline
\end{tabular}
\end{table}
\end{Verbatim}
Notice the \verb!\hline! commands at the beginning and end of the
\texttt{tabular} environment to generate the double lines at the top
and the bottom of the table, as well as the single \verb!\hline! command
to separate the heading from the data. The typeset result is shown next.
Table~\ref{tab:famous} shows three famous mathematical constants.
\begin{table}[h]
\caption{Famous constants}
\label{tab:famous}
\centering
\begin{tabular}{@{}lll@{}}\hline\hline
Symbol & Description & Approximate value\\\hline
$e$ & base of natural logarithm & $2.7182818285$ \\
$\pi$ & ratio circle circumference to diameter &
$3.1415926536$ \\
$\phi$ & golden ratio & $1.6180339887$ \\\hline\hline
\end{tabular}
\end{table}
\subsection{Bibliographical references}
\label{sec:biblioref}
References should be cited in the text using numbers within square
brackets. This is achieved with the \verb!\cite! command. Punctuation
can be used either within or outside the brackets, but the same method
should be used consistently throughout the contribution. A few examples follow.
\begin{Itemize}
\item \verb!Ref.~\cite{bib:mybiba}! typesets, \eg `Ref.~[1]'.
\item \verb!Refs~\cite{bib:mybiba}--\cite{tab:mybibb}!,
typesets, \eg `Refs~[1--5].
\item At the beginning of a sentence the word ``Reference'' should be
written out in full. For instance \verb!Reference~\cite{bib:mybibc}!
which gives, \eg `Reference~[3]'.
\end{Itemize}
The text of the document should show the bibliographic references in
\emph{consecutive numerical order}. References in tables should be
entered in the order left to right, and top to bottom.
\subsubsection{List of references}
The list of references (the \verb!\bibitem! entries) must all be
grouped inside a \texttt{thebibliography} environment, as follows.
\begin{Verbatim}[numbers=left,fontsize=\small]
\section{Section title}
Work by Einstein~\cite{bib:einstein} as well as that in
Ref.~\cite{bib:gravitation} explain the theory of relativity.
...
\section{Introduction}
The biological effect of radiation depends on the quality of
the radiation as well as on the amount of energy absorbed. Evidence
suggests that this quality dependence is primarily caused by
the differences in rates of energy loss~\cite{Raby1966,Dupont1961}.
The general criteria that have been used are:
\begin{enumerate}
\item[i)]
to terminate the iteration when the residue between iterated
and experiment values is of the order of experimental
errors~\cite{Raby1966,Appleman1959,vanBerg1965,Bryant1985,Allen1977};
\item[ii)]
to terminate when the smoothest solution has been obtained.
\end{enumerate}
For an overall appreciation of the work carried out in this field,
see Refs.~\cite{Keil1969} and \cite{Guignard1983}.
\section{Examples of figures and tables}
\subsection{Including tabular material}
Smith tabulated the average values of mean linear energy transfer
(LET) obtained by the different methods used in
Ref.~\cite{Appleman1959}, and these are reproduced in
Table~\ref{tab:LET}. Note that a table is produced with \LaTeX's
\texttt{table} environment, and that the caption should be positioned
\emph{above} the tabular material.
\begin{table}[!h]
\begin{center}
\caption{Calculated mean LET values in water (keV/mm)}
\label{tab:LET}
\begin{tabular}{|p{6cm}|c|c|}
\hline
Radiation & Smith$^{a)}$ & Jones\\
\hline
1 MeV \ensuremath{\alpha} & & \\
200 kVp X-rays total & 3.25 & 1.79 \\
\hline
200 kVp X-rays (primary) & 2.60 & 1.48 \\
\hline
\multicolumn{3}{@{}l}{$^{a)}$ \footnotesize J. Smith, Rad. Res. 1 (1956) 234}
\end{tabular}
\end{center}
\end{table}
Table~\ref{tab:LET} is reproduced from the publication mentioned
earlier and shows the good agreement between predictions and
calculations. Comparison should be made with the decay curves shown
in Figs.~6 and 7 of Ref.~\cite{vanBerg1965}, and further information
is given in Section~\ref{sec:curvature} and Appendix~\ref{sec:app}.
\subsection{Including figures}
Figures are defined with the \texttt{figure} environment. The source
for producing the figure can be included inline, as in
Fig.~\ref{fig:sincos} which shows how the $\sin$ and $\cos$ functions
evolve with the help of the \texttt{picture} environment and the
\texttt{pict2e} and \texttt{curve2e} packages.
\begin{figure}[ht]
\begin{center}
\setlength\unitlength{1cm}
\begin{picture}(6.6,3)(0,-1.5)
\put(0,-1.5){\vector(0,3){3}}
\put(0,0){\Vector(6.6,0)}
\put(6.3,0.1){$\theta$}
\put(0.05,-0.3){$0$
\put(2.9,-0.3){$\pi$
\put(6.23,-0.3){$2\pi$
\color{red}\put(0.1,1.1){$\cos\theta$}
\Curve(0,1)<1,0>(1.570796,0)<1,-1>%
(3.1415924,-1)<1,0>(6.283185,1)<1,0>%
\color{blue}\put(1.65,1.1){$\sin\theta$}
\Curve(0,0)<1,1>(1.570796,1)<1,0>%
(4.712389,-1)<1,0>(6.283185,0)<1,1>%
\end{picture}
\end{center}
\caption{The $\sin$ and $\cos$ functions}
\label{fig:sincos}
\end{figure}
Figures can also be imported, in EPS or PDF, PNG and JPEG format,
depending on whether you run \texttt{latex} or
\texttt{pdflatex}. Figure~\ref{fig:extpict} show how to include a
picture with the \verb|\includegraphics| command of \LaTeX's
\texttt{graphicx} package.
\begin{figure}[ht]
\centering
\includegraphics[width=.5\linewidth]{testpicture}
\caption{Including a figure from an external file}
\label{fig:extpict}
\end{figure}
\subsection{Examples of equations}
\label{sec:curvature}
Equation~\ref{eq:a1} representing a straight line at an angle
$\theta$, is
\begin{equation}
n^k(h)u=\lambda h t g q_k \label{eq:a1}
\end{equation}
and
\begin{equation}
n^k(h) =k h \frac{k}{32} \label{eq:a2}
\end{equation}
where:
\begin{itemize}
\item $\lambda$ is the distance between two consecutive sweep lines,
\item $u$ is the least count.
\end{itemize}
We consider a parabola, the tangent of which is parallel to the
vertical axis of the main matrix. A sufficiently good approximation
to a parabola [see Eq.~\ref{eq:a3}] drawn inside the matrix can be given
by the following formula:
\begin{equation}
n_q = \alpha q^2 \quad \alpha=\frac{\lambda^2}{3 Ru} \label{eq:a3}
\end{equation}
where $R$ is the radius of curvature. We have the following
relationship:
\begin{equation}
n_q + \mu_q \text{ with } \mu_q = \alpha (2q + 1). \label{eq:a4}
\end{equation}
\section{Conclusion}
The theoretical considerations presented have been confirmed
by their close agreement with the results of practical experiments.
An account of the earlier work carried out in this field can
be found in the bibliography.
It is expected that in the next few years many new results will
be published, since a significant number of new experiments have
recently been launched.
\section*{Acknowledgements}
We wish to thank C. Brown for his most enlightening comments
on this topic.
\section{Introduction}
The biological effect of radiation depends on the quality of
the radiation as well as on the amount of energy absorbed. Evidence
suggests that this quality dependence is primarily caused by
the differences in rates of energy loss~\cite{Raby1966,Dupont1961}.
The general criteria that have been used are:
\begin{enumerate}
\item[i)]
to terminate the iteration when the residue between iterated
and experiment values is of the order of experimental
errors~\cite{Raby1966,Appleman1959,vanBerg1965,Bryant1985,Allen1977};
\item[ii)]
to terminate when the smoothest solution has been obtained.
\end{enumerate}
For an overall appreciation of the work carried out in this field,
see Refs.~\cite{Keil1969} and \cite{Guignard1983}.
\section{Examples of figures and tables}
\subsection{Including tabular material}
Smith tabulated the average values of mean linear energy transfer
(LET) obtained by the different methods used in
Ref.~\cite{Appleman1959}, and these are reproduced in
Table~\ref{tab:LET}. Note that a table is produced with \LaTeX's
\texttt{table} environment, and that the caption should be positioned
\emph{above} the tabular material.
\begin{table}[!h]
\begin{center}
\caption{Calculated mean LET values in water (keV/mm)}
\label{tab:LET}
\begin{tabular}{|p{6cm}|c|c|}
\hline
Radiation & Smith$^{a)}$ & Jones\\
\hline
1 MeV \ensuremath{\alpha} & & \\
200 kVp X-rays total & 3.25 & 1.79 \\
\hline
200 kVp X-rays (primary) & 2.60 & 1.48 \\
\hline
\multicolumn{3}{@{}l}{$^{a)}$ \footnotesize J. Smith, Rad. Res. 1 (1956) 234}
\end{tabular}
\end{center}
\end{table}
Table~\ref{tab:LET} is reproduced from the publication mentioned
earlier and shows the good agreement between predictions and
calculations. Comparison should be made with the decay curves shown
in Figs.~6 and 7 of Ref.~\cite{vanBerg1965}, and further information
is given in Section~\ref{sec:curvature} and Appendix~\ref{sec:app}.
\subsection{Including figures}
Figures are defined with the \texttt{figure} environment. The source
for producing the figure can be included inline, as in
Fig.~\ref{fig:sincos} which shows how the $\sin$ and $\cos$ functions
evolve with the help of the \texttt{picture} environment and the
\texttt{pict2e} and \texttt{curve2e} packages.
\begin{figure}[ht]
\begin{center}
\setlength\unitlength{1cm}
\begin{picture}(6.6,3)(0,-1.5)
\put(0,-1.5){\vector(0,3){3}}
\put(0,0){\Vector(6.6,0)}
\put(6.3,0.1){$\theta$}
\put(0.05,-0.3){$0$
\put(2.9,-0.3){$\pi$
\put(6.23,-0.3){$2\pi$
\color{red}\put(0.1,1.1){$\cos\theta$}
\Curve(0,1)<1,0>(1.570796,0)<1,-1>%
(3.1415924,-1)<1,0>(6.283185,1)<1,0>%
\color{blue}\put(1.65,1.1){$\sin\theta$}
\Curve(0,0)<1,1>(1.570796,1)<1,0>%
(4.712389,-1)<1,0>(6.283185,0)<1,1>%
\end{picture}
\end{center}
\caption{The $\sin$ and $\cos$ functions}
\label{fig:sincos}
\end{figure}
Figures can also be imported, in EPS or PDF, PNG and JPEG format,
depending on whether you run \texttt{latex} or
\texttt{pdflatex}. Figure~\ref{fig:extpict} show how to include a
picture with the \verb|\includegraphics| command of \LaTeX's
\texttt{graphicx} package.
\begin{figure}[ht]
\centering
\includegraphics[width=.5\linewidth]{picture}
\caption{Including a figure from an external file}
\label{fig:extpict}
\end{figure}
\subsection{Examples of equations}
\label{sec:curvature}
Equation~\ref{eq:a1} representing a straight line at an angle
$\theta$, is
\begin{equation}
n^k(h)u=\lambda h t g q_k \label{eq:a1}
\end{equation}
and
\begin{equation}
n^k(h) =k h \frac{k}{32} \label{eq:a2}
\end{equation}
where:
\begin{itemize}
\item $\lambda$ is the distance between two consecutive sweep lines,
\item $u$ is the least count.
\end{itemize}
We consider a parabola, the tangent of which is parallel to the
vertical axis of the main matrix. A sufficiently good approximation
to a parabola [see Eq.~\ref{eq:a3}] drawn inside the matrix can be given
by the following formula:
\begin{equation}
n_q = \alpha q^2 \quad \alpha=\frac{\lambda^2}{3 Ru} \label{eq:a3}
\end{equation}
where $R$ is the radius of curvature. We have the following
relationship:
\begin{equation}
n_q + \mu_q \text{ with } \mu_q = \alpha (2q + 1). \label{eq:a4}
\end{equation}
\section{Conclusion}
The theoretical considerations presented have been confirmed
by their close agreement with the results of practical experiments.
An account of the earlier work carried out in this field can
be found in the bibliography.
It is expected that in the next few years many new results will
be published, since a significant number of new experiments have
recently been launched.
\section*{Acknowledgements}
We wish to thank C. Brown for his most enlightening comments
on this topic.
\section{Introduction}
Heavy-ion collisions at ultra-relativistic energies produce strongly
interacting matter under extreme conditions. The main goal is to
create in these reactions a state in which the confinement of quarks
and gluons inside hadrons is no longer effective, the so-called
quark-gluon plasma. This strongly compressed matter undergoes a rapid
expansion with a drop of temperature and energy density. Two-particle
momentum correlations provide unique information on the size and
dynamic evolution of this fireball and are therefore a widely employed
observable in heavy-ion physics. Usually correlations of identical
charged pions are studied which, due to the high available statistics,
allow a multi-dimensional study of radius parameters \cite{PRATT86,
BERTSCH89,NA49HBT}. Less frequently two-proton correlations are
analyzed \cite{NA49PP,NA44PP,WA80PP,E895PLAM}, although only
one-dimensionally. Moreover, the large abundance of strange particles
produced in heavy-ion collisions allows to study also two particle
correlations between pairs of strange particles or pairs of strange
and non-strange particles. For example the correlations of identical
kaons were investigated at the CERN-SPS \cite{NA49KK} as well as at
RHIC \cite{STARKK}. In this paper we report on the measurement of the
\lamp\ correlation function in central Pb+Pb collisions at \sqrts~=
17.3~GeV at the CERN-SPS.
It was suggested that also the momentum correlation between \lam\ and
protons can be employed to measure the size of the emitting source
\cite{WANG99}. The correlation function of \lamp\ pairs is only
affected by the strong interaction between the particles. This
distinguishes \lamp\ correlations from the two-proton case, for which
the correlation function is dominated by the repulsive Coulomb
interaction and the Fermi-Dirac statistics at low relative
momenta. Both effects are absent in the \lamp\ correlation function,
which should therefore be more sensitive to large source
sizes \cite{WANG99}. However, the knowledge of the strong
interaction between protons and \lam\ is necessary to relate the
strength of the correlation to the size of the emitting source. There
is a substantial set of data available on low-energy elastic $\lam$~p
scattering \cite{ALEXANDER68,SECHIZORN68,KADYK71}, and on
\mbox{$\kmin \textrm{d} \rightarrow \lam \: \textrm{p} \: \pimin$}
\cite{DAHL61,CLINE68,BRAUN77}, as well as
\mbox{$\textrm{p} \textrm{p} \rightarrow \textrm{p} \kplus \lam$}
\cite{COSY11,HIRES} reactions. Also \lam\ hypernuclei provide
important information on the \lam\ nucleon interaction. Based on this
data many theoretical analyses derived \lamp\ scattering lengths and
effective interaction ranges \cite{NAGELS79,BODMER88,MAESSEN89,
RIJKEN99,HAIDENBAUER05,POLINDER06,SIBIRTSEV06} which can be used to
calculate the \lamp\ correlation function.
A preliminary study by the NA49 experiment at the SPS was
reported in \cite{BLUME02}. Here we describe the final results of a
new and improved analysis \cite{BECK09}. Similar studies of \lamp\
correlations in heavy ion collisions were performed at lower
\cite{HADESPLAM,E895PLAM} and higher \cite{STARPLAM} center-of-mass
energies, allowing to investigate the evolution of \lamp\ correlations
with \sqrts.
\section{Data Analysis}
The data presented here were measured by the NA49 experiment at
the CERN SPS. A detailed description of the experimental setup can be
found in \cite{NA49NM}. Charged particles produced by interactions
of the Pb beam in a thin Pb-foil target are tracked with four
large-volume Time Projection Chambers (TPCs).
Two TPCs are placed inside two superconducting dipole
magnets, while the other two are situated outside of the magnetic
field. Since the latter measure long pieces of the particle tracks,
they allow a precise determination of the specific energy loss \dedx\
inside the detector gas (typical resolution of 4~\%) and thus particle
identification in a large region of phase-space. Additional particle
identification is provided around mid-rapidity by Time-Of-Flight
detectors. A Zero Degree Calorimeter is used to measure the energy in
the projectile fragmentation region from which the centrality of the
reaction can be deduced. This analysis is based on \mbox{$2.8 \cdot
10^{6}$} Pb+Pb events at \sqrts~= 17.3~GeV recorded in the year
2000, which cover the 23.5~\% most central part of the total inelastic
cross section, corresponding to an averaged number of wounded
nucleons of $\nwound = 262$.
\subsection{\lam\ reconstruction}
The \lam\ hyperons are detected via their charged decay
\mbox{$\Lambda \rightarrow \textrm{p} \pimin$}, using the same methods
as described in \cite{NA49EHYP,NA49SHYP}. The \lam\ reconstruction is
done by forming pairs of positively and negatively charged tracks and
extrapolating them towards the main interaction vertex. The
positively (negatively) charged tracks are required to have at least
50 (30) reconstructed points. Pairs with a distance of closest
approach of less than 0.5~cm anywhere between the position of the
first measured point on the tracks and the target plane are considered
as \vzero\ candidates. Assigning proton and pion masses to the
positively and negatively charged decay particle, the invariant mass
of a \lam\ candidate is calculated. A significant reduction of the
combinatorial background can be achieved by applying several selection
criteria to the \lam\ candidates. In this analysis it is required
that the secondary vertex is separated by at least 25~cm in
beam-($z$)-direction from the target plane. Additionally, the
back-extrapolation of the flight path of the \lam\ candidate must not
deviate from the interaction vertex position in the transverse
directions $x$ and $y$ by more than $|\Delta x| = 0.75$~cm and
$|\Delta y| = 0.375$~cm. The signal-to-background ratio is further
improved by enriching the protons in the sample of positively charged
tracks by applying a momentum-dependent cut on the measured energy
loss (\dedx). An additional \dedx\ cut on the negatively-charged
tracks also allows to reject electrons from photon conversions.
\begin{figure}[t]
\includegraphics[height=80mm]{lam_minv.eps}
\caption{\label{fig:lam_signal}
The invariant mass distribution of \lam\ candidates in two exemplary
phase-space bins ((a) $-1.2 < y < -0.4$ and $0.6 < \pt < 1.2$~\gevc,
(b) $-0.4 < y < 0.4$ and $1.8 < \pt < 2.4$~\gevc) for central Pb+Pb
reactions at \sqrts~= 17.3~GeV. The two vertical lines indicate the
mass windows used to define \lam\ candidates.}
\end{figure}
An important point with respect to correlation studies is the
requirement that each \lam\ candidate must be unique. If it happens
that a daughter track of a given \lam\ candidate is also assigned to
another one, a strong artificial correlation between both candidates
is created which in turn affects the measured \lamp\ correlation
function. In this analysis it is therefore ensured that any given
track is used only once as a daughter track. Similarly, \lam\
daughters that were not reconstructed as a single track, but as two
track pieces (split tracks), will cause a distortion of the measured
correlation function. To exclude these tracks it is required that the
number of measured points of each accepted track is higher than 50~\%
of the number of points that this track could maximally have according
to its trajectory in the TPCs.
Figure~\ref{fig:lam_signal} shows the distribution of invariant mass
$m_{inv}$ of \lam\ candidates obtained after assigning the proton
($\pi^{-}$) mass to the positive (negative) daughter track for two
intervals of the center-of-mass rapidity $y$ and transverse momentum
$p_t$. \lam\ are accepted in a mass window \mbox{$[\mzero - \Delta
m, \mzero + \Delta m]$} of a half-width of \mbox{$\Delta m$ =
2~\mevcc}, where \mbox{\mzero\ = 1.115683~\gevcc} is the literature
value for the \lam\ mass \cite{PDG08}. This mass window was chosen in
order to optimize the signal-to-background ratio and reduce the
corrections for the signal purity.
\begin{figure}[t]
\includegraphics[height=80mm]{purity_lm.eps}
\caption{\label{fig:purity}
The purity $P_{\Lambda}(y,\pt)$ of the selected \lam\ candidates as
function of rapidity $y$ and transverse momentum \pt.
}
\end{figure}
The ratio between signal and background inside this mass window and
with it the purity of the \lam\ sample vary over phase-space. A
$\chi^{2}$-fit to the invariant mass distributions in different $y$
and \pt\ bins was performed in order to determine the relative
contributions. The fit function used is the sum of a polynomial for the
combinatorial background and a function for the \lam\ signal. The
measured shape of the \lam\ signal results from a convolution of the
resolutions for the single track momenta and the secondary vertex
positions and is found to be well described by an asymmetric
Lorentz-curve. By subtracting the background function from the
measured invariant mass distribution the number of real \lam\ is
determined. Defining the purity $P_{\Lambda}$ by the ratio of the
number of real \lam\ to the number of accepted \lam\ candidates in the
chosen mass window one obtains the result shown in \Fi{fig:purity}.
The phase-space averaged value of the \lam\ purity is $\langle
P_{\Lambda} \rangle = 69$~\%.
\begin{figure}[t]
\includegraphics[height=80mm]{dedx_fits.eps}
\caption{\label{fig:dedx_fits}
The \dedx\ spectra in two exemplary total momentum bins ((a) $4.0 < p
< 5.0$~\gevc, (b) $12.6 < p < 15.9$~\gevc) together with their
decomposition into contributions from protons, kaons, pions, and
electrons (from left to right). Shown is the measured energy loss
normalized to the minimum ionizing value $\dedx_{\rb{MIP}}$. The
vertical lines represent the upper \dedx\ cuts that result in a purity
of $P_{\rb{p}}= 0.8$.
}
\end{figure}
\subsection{Proton reconstruction}
Protons are identified via their energy loss \dedx\ as measured in
the two Main-TPCs. A valid proton track is required to have at least
50 reconstructed points. An additional cut on the impact parameters
in the target plane ($|b_{\rb{x}}| < 5.0$~cm, $|b_{\rb{y}}| < 2.0$~cm)
reduces the contribution from secondary tracks. In order to assign the
probability of being a proton to a given track, the energy loss
spectra measured in bins of total momentum $p$ are fitted by a sum
$E(x,p)$ of asymmetric Gaussians using a \chisq\ minimizing
procedure~\cite{MILICA}:
\begin{equation}
E(x,p) = \sum\limits_{i=\rb{d,p,K,}\pi,\rb{e}} A_{i}(p) \:
\frac{1}{\sum_{l} n_{l}} \:
\sum\limits_{l} \frac{n_{l}}{\sqrt{2 \pi} \: \sigma_{i,l}(p)}
\exp \left[ - \frac{1}{2} \left(
\frac{x - \hat{x}_{i}(p)} {(1 \pm \delta)\:\sigma_{i,l}(p)}
\right)^{2} \right] .
\end{equation}
Here, $A_{i}(p)$ denotes the yield for particle type $i$, $n_{l}$ the
number of tracks in a given track length interval $l$,
$\hat{x}_{i}(p)$ the most probable \dedx\ values for particle type
$i$, $x$ the measured \dedx\ value of the track under consideration,
$\sigma_{i,l}(p)$ the width of the Gaussian, and $\delta$ the
asymmetry parameter. The parameters $A_{i}(p)$, $\hat{x}_{i}(p)$, and
$\sigma_{\pi}(p)$ are determined by the fitting procedure in each
$p$~bin separately. The widths for the other particles types and
different track length bins are derived from the parameter
$\sigma_{\pi}(p)$ for pions: $\sigma_{i,l}(p) = \sigma_{\pi}(p) \cdot
(\hat{x}_{i}/\hat{x}_{\pi})^{\alpha}(1/\sqrt{l})$. The exponent was
determined to be $\alpha = 0.625$ and the parameter $\delta$ is
fixed to 0.065. Results of the fit procedure for two different total
momentum bins are shown in \Fi{fig:dedx_fits}. Since for low particle
momenta the energy loss curves of different particle species cross
each other and particle identification is thus not possible, a lower
cut on the total momentum of $p > 4$~\gevc\ is applied. To exclude
the region of the Fermi plateau also momenta above 50~\gevc\ are
discarded. Based on these fits a momentum dependent cut on the
measured \dedx\ values for single tracks is defined such that the
accepted tracks always have the same probability of being a proton.
For the standard analysis this probability is set to 80~\%, equivalent
to a constant proton purity $P_{\rb{p}}= 0.8$. As in the case of the
\lam\ decay daughters it is ensured that split tracks are removed from
the track sample by rejecting tracks that have less than 55~\% of the
number of geometrically possible points.
\subsection{Determination of the \lamp\ correlation function}
The selected \lam\ and proton candidates are then combined to form
\lamp\ pairs. In order to avoid trivial auto-correlations a track
that is used to reconstruct the \lam\ candidate is removed from the
primary proton sample. The pair distribution $S(\qinv)$ is measured
as a function of the generalized invariant relative momentum of the
\lamp\ pair \qinv, which is defined as the modulus of \mbox{$\tilde{q}
= q - P (q P) / P^{2}$}, with $q = p_{\rb{p}} - p_{\lam}$, $P =
p_{\rb{p}} + p_{\lam}$, and $q P = m_{\rb{p}}^{2} - m_{\lam}^{2}$.
Here, $p_{\rb{p}}$ and $p_{\lam}$ are the 4-momenta of the proton and
the \lam. In the two particle center-of-mass system $\tilde{q}$
reduces to $\{0, 2 \vec{k}^{*}\}$, with $2 \vec{k}^{*}$ being the
3-momentum difference in this reference frame \cite{LEDNICKY05}.
An event-mixing method that combines proton and \lam\ candidates taken
from different events, is employed for the construction of the
uncorrelated background $B(\qinv)$. The measured correlation function
is thus defined as:
\begin{equation}
\label{eq:meascf}
C_{\rb{meas}}(\qinv) = N \: \frac{S(\qinv)}{B(\qinv)} .
\end{equation}
The normalization constant $N$ is determined by requiring
$C_{\rb{meas}}(\qinv) = 1$ in the region $0.2 < \qinv < 0.3$~GeV.
Since the reconstruction of real pairs is affected by the limited
two-track resolution of the detector \cite{NA49NM}, a distance cut
between the track of the primary proton and the track of the positive
\lam\ decay particle is applied for real pairs as well as for the
mixed event pairs. For each pair it is required that the tracks have
an average separation of at least 3.0~cm. This average is determined
as the arithmetic mean of the track distances determined in planes
perpendicular to the beam axis. For each TPC two planes are taken
into account. Their distances to the target plane are 160.5~cm,
240.5~cm, 540.0~cm, 620.0~cm, 910.0~cm, and 990.0~cm. Pairs that do
not pass the distance cut are discarded.
\begin{figure}[t]
\includegraphics[height=80mm]{phasespace_lm_p_pairs.eps}
\caption{\label{fig:acceptance}
The phase-space population of protons (a), $\Lambda$ (b) as a function
of rapidity $y$ and transverse momentum \pt. The distribution for the
resulting \lamp\ pairs (c) is shown versus pair rapidity
$y_{\rb{p}\lam} = \frac{1}{2} \ln \frac{E_{\rb{p}} + E_{\lam} +
p_{\rb{z,p}} + p_{\rb{z},\lam}}{E_{\rb{p}} + E_{\lam} - p_{\rb{z,p}} -
p_{\rb{z},\lam}}$ and $\kt = \frac{1}{2} | \vec{p}_{\rb{t,p}} +
\vec{p}_{\rb{t},\lam} |$.
}
\end{figure}
Figure~\ref{fig:acceptance} shows the phase-space population of the
accepted protons, \lam, and \lamp\ pairs. Averaging over all measured
\lamp\ pairs we find \mbox{$\langle \kt \rangle = 0.53$~\gevc} with
$\kt = \frac{1}{2} | \vec{p}_{\rb{t,p}} + \vec{p}_{\rb{t},\lam} |$
(\mbox{$\langle \mt \rangle = 1.18$~GeV} with $\mt = \sqrt{\kt^{2} +
(\frac{1}{2}(m_{\lam} + m_{\rb{p}}))^{2}}$).
The measured correlation function can be affected by the finite
momentum resolution of the detector. In \cite{NA49HBT} an extensive
investigation of its influence on the radius parameters extracted from
correlations of identical charged pions as measured with the NA49
experiment is discussed. Due to the excellent momentum resolution of
the NA49 detector, it turned out that the impact is negligible and a
correction for this effect is not necessary. Even though the momentum
resolution for a \lam\ ($\langle \sigma_{p_{\lam}} \rangle \leq 1 \%$)
is worse than for the primary track, the resulting effect
on the measured \lamp\ correlation function is still clearly smaller
than all other systematics effects. Therefore no correction is
applied in this analysis.
A substantial fraction of the measured protons and \lam\ originate
from weak and electro-magnetic decays of heavier particles
(feed-down). In the following it will be assumed that these decay
particles are not correlated, because the decays happen long after the
thermal freeze-out, and will thus reduce the observed \lamp\
correlation function. For the \lam\ the feed-down originates from
\xim, \xizero, and \sig\ decays, while in case of the protons the
decays of \lam, and \sip\ contribute. We calculate the fraction of
protons and \lam\ originating from feed-down ($F_{\rb{p}}(y,\pt)$ and
$F_{\Lambda}(y,\pt)$) via a simulation procedure. \lam, \xim, and
\xizero\ are generated according to their measured phase-space
distributions \cite{NA49EHYP}. The measured \lam\ include the \sig\
which cannot be separated experimentally. For the \xizero\ the same
input distributions are assumed as for the \xim\, scaled by the ratio
\xizero/\xim\ of the total multiplicities taken from statistical model
fits \cite{BECATTINI06}. The daughter tracks of the generated
particles are followed through the NA49 detector setup using the
Geant3.21 package \cite{GEANT}. The response of the TPCs to the
traversing particles is simulated with NA49 specific software. In a next
step the simulated raw signals are added to measured raw data and
processed by the same reconstruction program as used for the
experimental data. By applying the same cuts as in the normal
analysis, the feed-down contributions to the measured \lam\ and
protons are determined. The $y$ and \pt\ dependences of the
feed-down fractions $F_{\rb{p}}(y,\pt)$ and $F_{\Lambda}(y,\pt)$ are
summarized in \Fi{fig:feeddown}. The phase-space averaged values are
$\langle F_{\rb{p}} \rangle = 22$~\% and $\langle F_{\Lambda} \rangle
= 43$~\%.
The final \lamp\ correlation function $C_{\rb{corr}}(\qinv)$ results
from the measured $C_{\rb{meas}}(\qinv)$ by applying a combined
correction factor $\langle K(\qinv) \rangle$ for purity and feed-down.
This factor is determined by averaging the product
$P_{\rb{p}}(y_{\rb{p}},p_{\rb{t,p}}) \, P_{\lam}(y_{\lam}, p_{\rb{t},\lam}) \:
(1 - F_{\rb{p}}(y_{\rb{p}},p_{\rb{t,p}})) \, (1 - F_{\lam}(y_{\lam},
p_{\rb{t},\lam}))$ over all reconstructed \lamp\ pair combinations
falling into a given bin of relative momentum \qinv. The corrected
correlation function thus follows from:
\begin{equation}
\label{eq:corrcf}
C_{\rb{corr}}(\qinv) = \frac{C_{\rb{meas}}(\qinv) - 1}
{\langle K(\qinv) \rangle} \: + 1 .
\end{equation}
\begin{figure}[t]
\includegraphics[height=80mm]{feeddown_lm_p.eps}
\caption{\label{fig:feeddown}
The fraction of protons ($F_{\rb{p}}(y,\pt)$) (a) and of \lam\
($F_{\Lambda}(y,\pt)$) (b) originating from feed-down as a function of
$y$ and \pt.}
\end{figure}
\section{Results}
In total 70920 \lamp\ pair candidates, corresponding to 17520 real
\lamp\ pairs (i.e. after correcting for purities and feed-down), with
\mbox{$\qinv < 0.2$~GeV} were measured. Dividing the signal
distribution by the event-mixing background and correcting with the
purities and feed-down contribution according to
\Eqs{eq:meascf}{eq:corrcf} yields the final \lamp\ correlation
function as shown in \Fi{fig:plamcf}. The correlation function
exhibits a significant enhancement for small \qinv. Such a
correlation would be expected as the effect of the strong interaction
between the proton and the \lam.
\subsection{Fit to theoretical calculation}
\begin{figure}[t]
\includegraphics[height=90mm]{plam_corrf.eps}
\caption{\label{fig:plamcf}
The corrected \lamp\ correlation function for central Pb+Pb reactions
at \sqrts~= 17.3~GeV, shown as a function of the invariant relative
momentum \qinv. The data represent an average over the whole
acceptance of the NA49 experiment. The lines display results of the
fit with a theoretical correlation function
\cite{LEDNICKY82,LEDNICKY05} (see text for details). Only statistical
errors are shown.
}
\end{figure}
Since the shape of the momentum correlation function was shown by Wang
and Pratt to depend on the size of the emitting source \cite{WANG99},
it can be used to extract its radius. The necessary prerequisite is a
quantitative knowledge of the \lamp\ interaction. Here we use a
functional form of the theoretical correlation function $C_{\rb{th}}$
that is based on the model of Lednick\'{y} and Lyuboshitz
\cite{LEDNICKY82,LEDNICKY05}. It employs an effective range
approximation of the S-wave \lamp\ interaction. The source size is
required to be larger than the effective range of the interaction.
The strength of the interaction is defined by four parameters: the
effective ranges \dzeros\ (\dzerot) and the scattering lengths
\fzeros\ (\fzerot) for the singlet $S = 0$ (triplet $S = 1$) state.
In our fits we use values of \dzeros~=~2.92~fm, \dzerot~=~3.78~fm,
\fzeros~=~-2.88~fm, and \fzerot~=~-1.8~fm, as suggested in
\cite{WANG99}. Under the assumption of unpolarized particle
production, the relative contribution of pairs in the singlet and the
triplet state is 1~:~3. Furthermore, we use the same spherically
symmetric Gaussian spatial distribution $S(\vec{r})$, for both the
proton and the \lam\ source:
\begin{equation}
S(\vec{r}) = \exp \left(- \frac{x^{2} + y^{2} + z^{2}}
{2 \rgaus^{2}} \right) .
\end{equation}
\begin{figure}[t]
\includegraphics[height=120mm]{syst_checks.eps}
\caption{\label{fig:systematics}
The corrected \lamp\ correlation function $C_{\rb{corr}}(\qinv)$ for
central Pb+Pb reactions at \sqrts~= 17.3~GeV shown as a function of
the invariant relative momentum \qinv\ for different purities of
protons (a) and \lam\ (b), as well as different feed-down
contributions to protons (c) and \lam\ (d) (see text). The lines
display the results of fits with a theoretical correlation
function~\cite{LEDNICKY82,LEDNICKY05}.
}
\end{figure}
Then, the theoretical correlation function $C_{\rb{th}}$ assumes the
functional form as quoted in reference \cite{STARPLAM}. By fitting it
to the data one obtains the effective radius parameter \rgaus\ and an
additional parameter $\lambda$ that takes possible reductions of the
height of the correlation into account:
\begin{equation}
\label{eq:fitcf}
C_{\rb{fit}}(\qinv) = \lambda \: (C_{\rb{th}}(\qinv) - 1) + 1 .
\end{equation}
These reductions occur, if the correction for the particle purities
and the feed-down, as defined in \Eq{eq:corrcf}, are insufficient. If
both parameters are left to vary freely, the best fit is obtained for
$\rgaus = 2.70 \pm 0.60$(stat.)~fm and $\lambda = 0.77 \pm
0.38$(stat.) (green dashed line in \Fi{fig:plamcf}). The agreement
within errors of the fitted $\lambda$ value with unity underlines the
consistency of the correction procedure. This justifies to fix
$\lambda = 1$ and to use only \rgaus\ as a free parameter, which
reduces the resulting error. With this constraint we obtain
$\rgaus = 3.02 \pm 0.20$(stat.)~fm (blue solid line in
\Fi{fig:plamcf}).
\subsection{Systematic uncertainties}
Systematic errors arise both in the extraction of the correlation
function and in the theoretical model. The former uncertainties were
studied by making small changes in the analysis procedure. By varying
the corresponding cuts on the measured energy loss used to identify
the primary protons, as well as the decay protons of the \lam\
candidates, the particle purities $P_{\rb{p}}(p_{\rb{p}})$ and
$P_{\lam}(y_{\lam}, p_{\rb{t},\lam})$ can be varied to a certain extent.
This changes the measured correlation function $C_{\rb{meas}}(\qinv)$.
However, after applying the appropriate correction factor $\langle
K(\qinv) \rangle$ the same corrected correlation function should be
obtained. Figure~\ref{fig:systematics} shows comparisons of
$C_{\rb{corr}}(\qinv)$ for different proton (\Fi{fig:systematics},a))
and \lam\ purities (\Fi{fig:systematics},b)). Even though the
correction factor changes quite dramatically ($\approx 45$~\% in case
of the proton purity), the resulting correlation functions agree quite
well. The systematic error on the effective radius parameter \rgaus\
is derived by taking the maximal difference of \rgaus\ as obtained by
fits to the different correlation functions. It is found that it
changes between $2.91 - 3.28$~fm ($2.98 - 3.29$~fm), if the proton
(\lam) purity is varied between $50 - 90$~\% ($64 - 72$~\%).
A similar study is performed by varying the feed-down contribution by
changing the cuts on the impact parameter $b_{\rb{x}}$ and
$b_{\rb{y}}$ ($\Delta x$ and $\Delta y$) of the proton tracks
(\Fi{fig:systematics},c) and \lam\ candidates
(\Fi{fig:systematics},d). One finds that \rgaus\ changes maximally
between $3.02 - 3.13$~fm ($2.92 - 3.23$~fm), if the contributions to
the protons (\lam) from feed-down is varied between $17 - 25$~\% ($39
- 46$~\%).
The dependence of the radius parameter on the region in \qinv\ that is
used to determine the normalization constant $N$ (see \Eq{eq:meascf})
is investigated by varying the size and position of this region. It
is found that \rgaus\ changes by maximally $\pm$0.11~fm.
By taking the quadratic sum of the different contributions a total
systematic error on \rgaus\ of $+0.44$~fm and $-0.16$~fm is
estimated.
Another source of systematic uncertainty of the extracted radius
parameter arises from the limited precision of the knowledge of
the scattering lengths and effective ranges used in the calculation of
the theoretical correlation function. Therefore, the fits were
repeated with \dzeros, \dzerot, \fzeros, and \fzerot\ taken from
\cite{COSY11,HIRES,BODMER88,SIBIRTSEV06,RIJKEN99,HAIDENBAUER05,LEDNICKY05}.
The largest deviation from the \rgaus\ value obtained with the
standard parameter set \cite{WANG99} is observed when using the
parameters extracted by the COSY-11 collaboration \cite{COSY11}.
While in all other cases the difference is smaller than $\pm$0.1~fm,
it is found to be $+0.274$~fm for the COSY-11 parameters which is
still close to our statistical error. Therefore we conclude that the
choice of the parameters describing the \lamp\ interaction has a
negligible effect on the final result.
\subsection{Energy dependence of the effective radius parameter}
\begin{figure}[t]
\includegraphics[height=100mm]{r_vs_sqrts.eps}
\caption{\label{fig:rsqrts}
The effective radius parameter \rgaus\ for \lamp\ correlations
as a function of center of mass energy \sqrts. The figure also
includes data on Ar+KCl collisions by the HADES collaboration
\cite{HADESPLAM} and on central Au+Au collisions by the E895
\cite{E895PLAM} and the STAR \cite{STARPLAM} experiments. The
systematic errors are represented by the brackets. Please note that
the STAR result corresponds to a slightly higher pair \mtavg\ than for
the NA49 measurement. Assuming the \mt~dependence shown
\Fi{fig:rmeanmt} the STAR data point would move up by $\approx 6$~\%,
if it was measured at the same \mtavg.
}
\end{figure}
The result of this study is compared to data at lower and higher
center-of-mass energies in \Fi{fig:rsqrts}. Good agreement with the
effective radius parameters measured for central Au+Au collisions by
the E895 collaboration at \sqrts~= 3.83~GeV \cite{E895PLAM} and by the
STAR experiment at \sqrts~= 200~GeV \cite{STARPLAM} is observed,
indicating that there is no significant change from AGS to RHIC
energies. A similar observation was made for correlations of
identical charged pions \cite{NA49HBT}.
\begin{figure}[t]
\includegraphics[height=100mm]{r_vs_meanmt_rsrorl.eps}
\caption{\label{fig:rmeanmt}
The effective radius parameter \rgaus\ extracted from the correlation
functions of \pimin\pimin\ \cite{NA49HBT}, \kplus\kplus\ and
\kmin\kmin\ \cite{NA49KK}, pp \cite{NA49PP}, and \lamp\ pairs versus
\mtavg\ as measured by NA49 for central Pb+Pb collisions at \sqrts~=
17.3~GeV. The systematic errors are represented by the brackets. For
explanation of curves see text.
}
\end{figure}
At even lower energies (\sqrts~= 2.61~GeV) the HADES collaboration
measured a significantly smaller effective radius parameter for \lamp\
correlations of \mbox{$\rgaus = 2.09 \pm 0.16$(stat.)~fm}
\cite{HADESPLAM}. However, the fireball volume in the case of Ar+KCl
collisions is expected to be smaller than for Pb+Pb and Au+Au
collisions. In fact, it was demonstrated in \cite{HADESPLAM}
that the measured \lamp\ radius parameter is dominated by the
reaction geometry and scales approximately as $A^{1/3}$, irrespective
of center-of-mass energy. A comparable observation was previously
made for two-proton correlations in the target fragmentation
region \cite{WA80PP}.
\subsection{\mtavg\ dependence of the effective radius parameter}
Figure~\ref{fig:rmeanmt} shows a comparison of \rgaus, as determined
in this analysis from the \lamp\ correlation function, to radius
parameters derived from correlations of \pimin\ pairs \cite{NA49HBT}
and charged kaon pairs \cite{NA49KK}, as well as from two-proton
correlations \cite{NA49PP}, at different average transverse masses
\mtavg. In case of the \pimin\pimin\ and KK correlations, \rgaus\
was calculated from the three-dimensional radius parameter components
as:
\begin{equation}
\label{eq:rzero}
\mbox{$\rgaus = (\rside \, \rout \, \rlong)^{1/3}$} .
\end{equation}
The decrease of the effective radius parameters with increasing
transverse mass is generally attributed to the presence of collective
flow in the fireball. The measurement of the \lamp\ correlation
allows to extend this study to higher \mtavg. In fact, \rgaus\ for
\lamp\ pairs is significantly smaller than the effective radius
parameter extracted for pions and kaons at lower \mtavg\ and is thus
in agreement with the expected behavior. The dashed curve in
\Fi{fig:rmeanmt} corresponds to a simple $\propto \mtavg^{-1/2}$
dependence. The solid line is based on the following \mt~dependences
of the three radius components, as suggested by hydrodynamical
approaches \cite{WIEDEMANN99,CHAPMAN95,MAKHLIN88}:
\begin{eqnarray}
\rside^{2} & = & \rgeo^{2} / (1 + (\mt/T) \: \etaf^{2})
\label{eq:rside} \\
\rout^{2} & = & \rside^{2} + \betat^{2} \: \Delta\tau^{2} \\
\rlong^{2} & = & \tauzero^{2} \: (T/\mt).
\label{eq:rlong}
\end{eqnarray}
Here \rgeo\ is the transverse size of the particle source, $\betat =
v_{\rb{f}}/c$ the transverse flow velocity, \etaf\ the transverse
flow rapidity $\etaf = (1/2) \: \log (1 + v_{\rb{f}})/(1 -
v_{\rb{f}})$, $T$ the kinetic freeze-out temperature, \tauzero\ the
total lifetime of the source and $\Delta\tau$ the emission duration.
Under the assumption that all particle species freeze out from the
same expanding source, we calculate \rgaus\ from \Eq{eq:rzero}, using
\Eqr{eq:rside}{eq:rlong}. The solid curve in \Fi{fig:rmeanmt}
corresponds to the parameter \etaf~=~0.8, \tauzero~=~0.6~fm/$c$,
$\Delta\tau$~=~3.4~fm/$c$ and $T$~=~90~MeV, as extracted by fits with
a blast wave model to the pion correlations \cite{NA49HBT}. The
overall normalization has been adjusted to fit the data. A reasonable
description of the effective radius parameters for most particle
species can thus be achieved, with the notable exception of the
two-proton correlation.
\section{Summary}
We report on the measurement of the \lamp\ correlation
function in momentum space for central Pb+Pb collisions at \sqrts~=
17.3~GeV. The \lamp\ pairs exhibit a clear positive correlation for
small relative momenta. By comparison to a calculated correlation
function a one-dimensional Gaussian source size parameter of
\mbox{$\rgaus = 3.02 \pm 0.20$(stat.)$^{+0.44}_{-0.16}$(syst.)~fm} is
determined. This value is in good agreement with measurements for
Au+Au collisions at lower and higher center-of-mass energies. The
\mtavg~dependence of the effective \lamp\ radius parameter follows the
expectation for an expanding source as described by hydrodynamics,
when compared to other two-particle correlation results.
\section*{Acknowledgments}
This work was supported by
the US Department of Energy Grant DE-FG03-97ER41020/A000,
the Bundesministerium fur Bildung und Forschung, Germany (06F~137),
the German Research Foundation (grant GA 1480/2-1),
the Polish Ministry of Science and Higher Education (1~P03B~006~30,
1~P03B~127~30, 0297/B/H03/2007/33, N~N202~078735, N~N202~078738,
N~N202~204638),
the Hungarian Scientific Research Foundation (T068506),
the Bulgarian National Science Fund (Ph-09/05),
the Croatian Ministry of Science, Education and Sport (Project
098-0982887-2878) and
Stichting FOM, the Netherlands.
|
\section{Conclusion}
\label{sec:concl}
We have described and measured the memory topology of two different high-end
machines using Intel and AMD processors.
These measurements demonstrate that NUMA effects exist and require engineering
beyond that normally employed to achieve good locality and cache use.
Further, we have shown that the NUMA penalty is significantly lower on Intel
systems due to the larger cross-processor bandwidth provided by QPI.
But, the AMD system provides greater total bandwidth for NUMA-aware
applications.
\paragraph{Acknowledgments}
Thanks to Bradford Beckmann for reviewing the breakdown of the AMD G34 socket.
The AMD machine used for the benchmarks was supported by National Science
Foundation Grant CCF-1010568 and this work is additionally supported in part by
National Science Foundation Grant CCF-0811389.
The views and conclusions contained herein are those of the authors and should
not be interpreted as necessarily representing the official policies or
endorsements, either expressed or implied, of these organizations or the
U.S.\ Government.
Access to the Intel machine was provided by Intel Research.
Thanks to the management, staff, and facilities of the Intel Manycore Testing
Lab.\footnote{Manycore Testing Lab Home:\\
\url{http://www.intel.com/software/manycoretestinglab}\\
Intel Software Network:\\
\url{http://www.intel.com/software}}
\section{Evaluation}
\label{sec:evaluation}
Our AMD test machine is described in \secref{intro:amdhardware}.
This machine runs x86\_64 Ubuntu Linux 10.04.2 LTS, kernel version 2.6.32-27.
\secref{intro:intelhardware} describes our Intel test machine.
This machine runs x86\_64 RedHat Enterprise Linux, kernel version
2.6.18-194.11.4.el5.
The vproc{} local heap size is 1~MB, which is slightly more than the L3~cache
size when nodes each have 6~threads.
We ran each experiment 10 times and we report the average performance results in
our graphs and tables.
\subsection{Benchmarks}
For our empirical evaluation, we use five benchmark programs from our benchmark suite
and one synthetic benchmark.
Each benchmark is written in a pure, functional style and was originally written
by other researchers and ported to our system.
The Barnes-Hut benchmark~\cite{barnes-hut} is a classic N-body problem solver.
Each iteration has two phases.
In the first phase, a quadtree is constructed from a sequence of mass points.
The second phase then uses this tree to accelerate the computation of
the gravitational force on the bodies in the system.
Our benchmark runs 20 iterations over 400,000 particles generated in
a random Plummer distribution.
Our version is a translation of a Haskell
program~\cite{barnes-hut-haskell-bench}.
The Raytracer benchmark renders a $512 \times 512$ image in parallel as
two-dimensional sequence, which is then written to a file.
The original program was written in ID~\cite{id90-manual} and is a simple
ray tracer that does not use any acceleration data structures.
The sequential version differs from the parallel code in that it
outputs each pixel to the image file as it is computed, instead of building
an intermediate data structure.
The Quicksort benchmark sorts a sequence of 10,000,000 integers in parallel.
This code is based on the {\textsc{Nesl}}{} version of the algorithm~\cite{scandal-algorithms}.
The SMVM benchmark is a sparse-matrix by dense-vector multiplication.
The matrix contains 1,091,362 elements and the vector 16,614.
The DMM benchmark is a dense-matrix by dense-matrix multiplication in which
each matrix is $600 \times 600$.
\begin{figure*}
\vspace*{-0.5in}
\begin{center}
AMD Speedups\\
\includegraphics[width=6in]{data/results/amd/speedup} \\[3em]
Intel Speedups\\
\includegraphics[width=6in]{data/results/intel/speedup} \\
\end{center}%
\caption{
Comparative speedup plots for five benchmarks on both AMD and Intel hardware.
The baseline is the single-processor version of each benchmark.
Larger speedup values are better, and speedup values equal to the number of
threads are ideal.
}
\label{fig:speedups}
\end{figure*}%
\subsection{Performance}
The dense-matrix multiplication (DMM) and raytracer benchmarks have abundant,
independent parallelism and our compiler and runtime exploit them, demonstrating
nearly ideal speedup up to the maximum number of cores on both machines.
Quicksort, sparse-matrix multiplication, and barnes-hut have excellent behavior
until either 24~or 36~threads, but then taper off.
On the AMD machine, both quicksort and barnes-hut scale nicely to 36~threads but
then only take slight advantage of additional threads.
In barnes-hut, we believe that this behavior is due to the sequential portion.
Quicksort also is limited by its fork-join parallelism, and without
significantly increasing the size of the underlying dataset, it is difficult to
take advantage of the additional available parallelism.
Sparse-matrix multiplication provides the least scalability for the AMD system.
We believe that this is due to a large amount of available execution parallelism
but a relatively small amount of data.
Unless this data is either perfectly divided between the nodes or replicated to
each location, this benchmark fails to take much advantage of greater than even
24~threads.
On the Intel machine, quicksort, barnes-hut, and spare-matrix multiplication
(SMVM) all see reducing speedups past 16~threads, but do not experience the more
serious lack of speedups seen on the AMD machine.
We believe that this better performance, particularly on SMVM, is due to a
smaller NUMA penalty when accessing the relatively smaller amount of shared
data, much of which resides on only one node.
Additionally, with only 4~nodes on the Intel machine, threads are twice as
likely to be located near data even if that data was placed randomly.
\paragraph{Locality affects scalability}
As these benchmarks and the figures in \secref{sec:numa} have shown, locality
and NUMA effects have an impact on benchmarks.
Benchmarks such as dense-matrix multiplication and raytracer, with excellent
locality and almost no shared data can scale nearly perfectly if all of their
data is kept locally.
The other benchmarks, which feature either heavily shared data or significant
points that sequentially merge data before creating more parallel work show
diminished improvements.
Unfortunately, these effects do not even easily show up on machines with
multiple processor packages until relatively large numbers of cores --- in our
experience, between 24~and~36.
Further, testing a runtime on a machine with a large number of cores but where
all are on the same processor package provides no feedback on the NUMA
scalability of that runtime.
\subsection{Load-balanced global collections}
\label{sec:evalGlobal}
During our global copying heap collection, there is often imbalanced available
work.
This imbalance results from uneven allocation patterns by threads,\footnote{In
particular, when a large data file is sequentially read into memory.} and
causes there to be extra blocks of to-space available to scan.
Prior work on the Glasgow Haskell Compiler (GHC) showed increased times when
parallel collections also performed global load-balancing of this imbalanced
work~\cite{multicore-haskell}.
As we show in \tblref{eval:load-balanced}, load-balancing is very effective when
performed on a per-node basis.
We compare the performance of a unbalanced collections with balanced
collections on our AMD machine, where in the latter case threads will scan
unscanned chunks generated from any thread, so long as it is on the same node.
Execution times are measured in seconds, and lower numbers are better.
The DMM, SMVM, and raytracer benchmarks are not included because none of those
programs keep data around for long enough to trigger a global GC collection.
Barnes-hut and quicksort both see greater than a 15\% reduction in global GC
time and a nearly 3\% reduction in overall execution time, when run on
48~cores.
\begin{table}
\begin{center}
\begin{tabular}{r | c | c | c | c}
& \multicolumn{2}{c|}{Unbalanced} & \multicolumn{2}{c}{Balanced} \\
Benchmark & Global (s) & Total (s) & Global (s) & Total (s)\\
\hline
Barnes-hut & 0.308 & 2.52 & 0.255 & 2.45 \\
Quicksort & 0.321 & 2.16 & 0.268 & 2.10 \\
\end{tabular}
\end{center}
\caption{
Comparison of load-balanced versus unbalanced global collections on 48
cores on the AMD machine. Smaller numbers are better.
}
\label{eval:load-balanced}
\end{table}%
\subsection{Single-thread performance versus a sequential program}
As is shown in \tblref{eval:mlton}, only on the raytracer do we offer
competitive performance to the sequential MLton baseline.
On all other benchmarks, we generally reach speed parity with four threads on
the AMD machine.
This performance gap is due to missed opportunities for sequential optimization
in Manticore and some small overhead from our parallel language constructs.
But, given that MLton has state-of-the-art functional language performance, this
comparison demonstrates that we have performance comparable with mainstream
functional languages on these benchmarks.
\begin{table}
\begin{center}
\begin{tabular}{r | c | c | c | c}
Benchmark & MLton (s) & 1T (s) & 2T (s) & 4T (s)\\
\hline
Barnes-hut & 20.4 & 61.4 & 38.2 & 17.0 \\
DMM & 12.4 & 50.6 & 27.6 & 12.6 \\
Quicksort & 20.9 & 100.8 & 52.5 & 27.5 \\
Raytracer & 10.9 & 15.6 & 7.7 & 3.9 \\
SMVM & 7.2 & 28.3 & 15.0 & 7.6\\
\end{tabular}
\end{center}
\caption{
Comparison of Manticore performance versus MLton at low numbers of
threads on the AMD machine. Smaller execution times are better.
}
\label{eval:mlton}
\end{table}%
\section{Introduction}
Inexpensive multicore processors and accessible multiprocessor motherboards
have brought all of the challenges inherent in parallel programming with large numbers of
threads with non-uniform memory access (NUMA) into the foreground.
Functional programming languages are a particularly interesting approach
to programming parallel systems, since they provide a high-level programming
model that avoids many of the pitfalls of imperative parallel programming.
But while functional languages may seem like a better fit for parallelism due to
their ability to compute independently while avoiding race conditions and
locality issues with shared memory mutation, implementing a scalable functional
parallel programming language is still challenging.
Since functional languages are value-oriented, their performance is highly
dependent upon their memory system.
This system is often the major limiting to improved performance in these
systems~\cite{multicore-haskell,intel-private-heap}.
Our group has been working on the design and implementation of a
parallel functional language to address the opportunity afforded by
multicore processors.
In this paper, we describe some benchmarks we have used to measure top-end
AMD and Intel machines to assist in the design and tuning of our parallel
garbage collector.
This paper makes the following contributions:
\begin{enumerate}
\item
We describe the architecture of and concretely measure the bandwidth and
latency due to the memory topology in both a 48-core AMD Opteron server and
a 32-core Intel Xeon server.
Looking only at technical documents, it is difficult to understand how much
bandwidth is achievable from realistic programs and how the latency of
memory access changes with increased bus saturation.
\end{enumerate}%
\subsection{AMD Hardware}
\label{intro:amdhardware}
Our AMD benchmark machine is a Dell PowerEdge R815 server, outfitted
with 48~cores and 128~GB physical memory.
The 48~cores are provided by four AMD Opteron 6172 ``Magny Cours'' processors~\cite{magny-cours,opteron},
each of which fits into a single G34 socket.
Each processor contains two nodes, and each node has six cores.
The 128~GB physical memory is provided by thirty-two 4~GB dual ranked RDIMMs,
evenly distributed among four sets of eight sockets, with one set for each processor.
As shown in \figref{amd:magny}, these nodes, processors, and RAM chips form a
hierarchy with significant differences in available memory bandwidth and number
of hops required, depending upon the source processor core and the target physical memory location.
Each 6~core node (die) has a dual-channel double data rate 3 (DDR3) memory
configuration running at 1333~MHz from its private memory controller to its own
memory bank.
There are two of these nodes in each processor package.
This processor topology is also laid out in \tblref{amd:topology}.
\begin{figure}
\centering
\includegraphics[scale=0.7]{pictures/G34}
\caption{
Interconnects for one processor in a quad AMD Opteron machine.
}
\label{amd:magny}
\end{figure}%
\begin{table}
\begin{center}
\begin{tabular}{r | c | c}
Component & Hierarchy & \# Total\\
\hline
Processor & 4 per machine & 4\\
Node & 2 per processor & 8\\
Core & 6 per node & 48 \\
\end{tabular}
\end{center}
\caption{
Processor topology of the AMD machine.
}
\label{amd:topology}
\end{table}%
Bandwidth between each of the nodes and I/O devices is provided by four 16-bit
HyperTransport 3 (HT3) ports, which can each be separated into two 8-bit HT3
links.
Each 8-bit HT3 link has 6.4~GB/s of bandwidth.
The two nodes within a package are configured with a full 16-bit link and an
extra 8-bit link connecting them.
Three 8-bit links connect each node to the other three packages in this four
package configuration.
The remaining 16-bit link is used for I/O.
\figref{numa:amdbandwidth} shows the bandwidth available between the different
elements in the hierarchy.
\begin{table}
\begin{center}
\begin{tabular}{r | c }
\multicolumn{1}{c|}{} & Bandwidth (GB/s)\\
\hline
Local Memory & 21.3 \\
Node in same package & 19.2 \\
Node on another package & 6.4 \\
\end{tabular}
\end{center}
\caption{
Theoretical bandwidth available between a single node (6~cores) and the rest of an AMD Opteron 4P system.
}
\label{numa:amdbandwidth}
\end{table}%
Each core operates at 2.1~GHz and has 64~KB each of instruction and data L1 cache and 512~KB of L2 cache.
Each node has 6~MB of L3 cache physically present, but, by default, 1~MB is
reserved to speed up cross-node cache probes.
\subsection{Intel Hardware}
\label{intro:intelhardware}
The Intel benchmark machine is a QSSC-S4R server with 32~cores and 256~GB
physical memory.
The 32~cores are provided by four Intel Xeon X7560 processors~\cite{xeon,qssc}.
Each processor contains 8~cores, which can be but are not configured to run with
2~simultaneous multithreads (SMT).
This topology is laid out in \tblref{intel:topology}.
As shown in \figref{intel:xeon}, these nodes, processors, and RAM chips form a
hierarchy, but this hierarchy is more uniform than that of the AMD machine.
\begin{figure}
\centering
\includegraphics[scale=0.7]{pictures/Xeon}
\caption{
Interconnects for one processor in a quad Intel Xeon machine.
}
\label{intel:xeon}
\end{figure}%
\begin{table}
\begin{center}
\begin{tabular}{r | c | c}
Component & Hierarchy & \# Total\\
\hline
Processor & 4 per machine & 4\\
Node & 1 per processor & 4\\
Core & 8 per node & 32\\
\end{tabular}
\end{center}
\caption{
Processor topology of the Intel machine.
}
\label{intel:topology}
\end{table}%
Each of the nodes is connected to two memory risers, each of which has a
dual-channel DDR3 1066~MHz connection.
The 4~nodes are fully connected by full-width Intel QuickPath Interconnect (QPI)
links.
\figref{numa:intelbandwidth} shows the bandwidth available between the different
elements in the hierarchy.
\begin{table}
\begin{center}
\begin{tabular}{r | c }
\multicolumn{1}{c|}{} & Bandwidth (GB/s)\\
\hline
Local Memory & 17.1 \\
Other Node & 25.6 \\
\end{tabular}
\end{center}
\caption{
Theoretical bandwidth available between a single node (8~cores) and the rest of an Intel Xeon system.
}
\label{numa:intelbandwidth}
\end{table}%
Each core operates at 2.266~GHz and 32~KB each of instruction and data L1 cache
and 256~KB of L2 cache.
Each node has 24~MB of L3 cache physically present but, by default, 3~MB is
reserved to speed up both cross-node and cross-core caching.
\section{Measuring NUMA effects}
\label{sec:numa}
In \secref{intro:amdhardware}, we described the exact hardware configuration and
memory topology of our 48~core AMD Opteron system.
We also described the 32~core Intel Xeon system in
\secref{intro:intelhardware}.
These systems are the subject of the NUMA tests below.
\subsection{STREAM benchmark}
The C language STREAM benchmark~\cite{stream} consists of the four operations
listed in \tblref{numa:stream}.
These synthetic memory bandwidth tests were originally selected to measure
throughput rates for a set of common operations that had significantly different
performance characteristics on vector machines of the time.
On modern hardware, each of these tests achieve similar bandwidth, as memory is
the primary constraint, not floating-point execution.
The \kw{COPY} test, in particular, is representative of the type of work
performed by a copying garbage collector.
\begin{table}
\begin{center}
\begin{tabular}{r | l}
Name & Code\\
\hline
\kw{COPY} & \lstinline!a[i] = b[i];! \\
\kw{SCALE} & \lstinline!a[i] = s*b[i];! \\
\kw{SUM} & \lstinline!a[i] = b[i]+c[i];! \\
\kw{TRIAD} & \lstinline!a[i] = b[i]+s*c[i];! \\
\end{tabular}
\end{center}
\caption{
Basic operations in the STREAM benchmark.
}
\label{numa:stream}
\end{table}%
The existing STREAM benchmark does not support NUMA awareness for either the
location of the running code or the location of the allocated memory.
We modified the STREAM benchmark to measure the achievable memory bandwidth for
these operations across several allocation and access configurations.
The baseline STREAM benchmark allocated a large, static vector of \kw{double}
values.\footnote{There is no difference in bandwidth when using \kw{long}
values.}
Our modifications use pthreads and libnuma to control the number and placement
of each piece of running code and corresponding
memory~\cite{butenhof:pthreads-book,libnuma}.
While the STREAM benchmark's suggested array sizes for each processor are larger
than the L3 cache, the tests do not take into account cache block sizes.
We extended the tests with support for strided accesses to provide a measure of
RAM bandwidth in the case of frequent cache misses.
This strided access support also allows us to measure the latency of memory
access.
\subsection{Bandwidth evaluation}
\figref{numa:evalBandwidth} plots the bandwidth, in MB/s, versus the number of
threads.
Larger bandwidth is better.
The results are for the \kw{COPY} test, but all of the tests were within a small
factor.
Four variants of the STREAM benchmarks were used:
\begin{enumerate}
\item \emph{Unstrided} accesses memory attached to its own node and uses the
baseline STREAM strategy of sequential access through the array.
\item \emph{Unstrided+non-NUMA} accesses the array sequentially, but is
guaranteed to access that memory on another package.
\item \emph{Strided} also accesses local memory, but ensures that each access is
to a new cache block.
\item \emph{Strided+non-NUMA} strides accesses, and also references memory from
another package.
\end{enumerate}
NUMA aware versions ensure that accessed memory is allocated on the same node as
the thread of execution and that the thread is pinned to the node, using the
libnuma library~\cite{libnuma}.
To do this, the modified benchmark pins the thread to a particular node and then
uses the libnuma allocation API to guarantee that the memory is allocated on the
same node.
The non-NUMA aware versions also pin each thread to a particular node, but then
explicitly allocate memory using libnuma from an entirely separate package (not
just a separate node on the same package).
When there are less threads than cores, we pin threads to new nodes rather than
densely packing a single node.
\begin{figure}
\centering
AMD Bandwidth\\
\includegraphics[scale=0.9]{data/amd/bandwidth} \\[3em]
Intel Bandwidth\\
\includegraphics[scale=0.9]{data/intel/bandwidth}
\caption{
Bandwidth vs. number of threads on the STREAM benchmark, comparing strided and NUMA configurations.
Larger bandwidth is better.
}
\label{numa:evalBandwidth}
\end{figure}%
It should not be surprising that the unstrided variants exhibit roughly eight
times the bandwidth of their strided versions, as cache blocks on these machines
are 64~bytes and the \kw{double} values accessed are each 8~bytes.
In the NUMA aware cases, scaling continues almost linearly until eight threads
and increases until the maximum number of available cores on both machines.
On AMD hardware, non-NUMA aware code pays a significant penalty and begins to
lose bandwidth where NUMA aware code does not at 48~cores.
On the Intel hardware the gap between NUMA and non-NUMA aware code is very small
even when the number of threads is the same as the number of cores.
But, the Intel hardware does not offer as much peak usable bandwidth for
NUMA-aware code, peaking near 40,000~MB/s whereas we achieve nearly 55,000~MB/s
on the AMD hardware.
\begin{figure}
\centering
AMD Bandwidth\\
\includegraphics[scale=0.9]{data/amd/bandwidth_by_node} \\[3em]
Intel Bandwidth\\
\includegraphics[scale=0.9]{data/intel/bandwidth_by_node}
\caption{
Bandwidth per node vs. number of threads on the STREAM benchmark, comparing
strided configurations.
Larger bandwidth is better.
}
\label{numa:evalBandwidthByNode}
\end{figure}%
\figref{numa:evalBandwidthByNode} plots the bandwidth against threads again,
but this time divided by the number of active nodes to provide a usage data
relative to the theoretical interconnect bandwidth detailed in
\tblref{numa:amdbandwidth} for the AMD machine and \tblref{numa:intelbandwidth}
for the Intel machine.
Our benchmarks allocate threads sparsely on the nodes.
Therefore, when there are less than 8~threads on the AMD machine, that is also
the number of active nodes.
On the Intel machine with 4~nodes, when there are less than 4~threads, that is
the number of active nodes.
These graphs show that on both machines there is a significant gap between
the theoretical bandwidth and that achieved by the strided \kw{COPY} stream
benchmark.
It is also clear that there is a significant non-NUMA awareness penalty on the
AMD machine but that penalty is less on the Intel machine.
However, the Intel machine begins to reach saturation at 32~cores, whereas the
NUMA aware AMD machine continues to increase per-node bandwidth up to 48~cores.
\begin{figure}
\centering
AMD Latencies\\
\includegraphics[scale=0.9]{data/amd/latency} \\[3em]
Intel Latencies\\
\includegraphics[scale=0.9]{data/intel/latency}
\caption{
Latency times versus the number of threads on the STREAM benchmark, comparing strided configurations.
Smaller latency times are better.
}
\label{numa:evalLatency}
\end{figure}%
\subsection{Latency evaluation}
\figref{numa:evalLatency} plots the latency times, in nanoseconds, versus the
number of threads.
Smaller latency times are better.
To measure the latency times, we only consider the strided STREAM benchmarks
to ensure that we are measuring only the time to access RAM, and not the time
to access cache.
As was the case with bandwidth, the AMD machine's NUMA aware tests maintain good
values up to large numbers of processors.
The non-NUMA aware AMD benchmark begins to exhibit high latencies at moderate
numbers of threads.
On the Intel machine, latency numbers remain low until more than 24~cores are in
use, and then the latencies grow similarly for both NUMA aware and non-NUMA
aware code.
On the AMD machine, these benchmarks and evaluations clearly indicate that at
high numbers of threads of executions, poor choices of memory location or code
execution can have significant negative impact on the latency of memory access
and memory bandwidth.
For the Intel machine, memory performance uniformly increases until high numbers
of threads, at which point is uniformly decreases, seeing little effect from
NUMA awareness.
But, none of these effects show up in practice until more than 24~threads are in
use.
|
\section{Computational Details}
We perform density functional theory calculations under the local density approximation as implemented in the AIMPRO code \cite{AIMPRO, AIMPRO2}. The calculations were carried out using supercells, fitting the charge density to plane waves within an energy cut-off of 300 Ha. Electronic level occupation was obtained using a Fermi occupation function with $kT$ = 0.04~eV. Relativistic pseudo-potentials are generated using the Hartwingster-Goedecker-Hutter scheme \cite{HGH}. These functions are labelled by multiple orbital symbols, where each symbol represents a Gaussian function multiplied by polynomial functions including all angular momenta up to maxima $p$ ($l$ = 0, 1) and $d$ ($l$ = 0, 1, 2). Following this nomenclature, the basis sets used for each atom type were $pdpp$ (B) and $ppp$ (H) (a more detailed account of the basis functions can be found elsewhere \cite{Goss2007}). A Bloch sum of these functions is performed over the lattice vectors to satisfy the periodic boundary conditions of the supercell. Structures were geometrically optimized with a single k-point for the B$_{12}$ cluster in a large 23.81 {\AA} cubic supercell, and a 4$\times$4$\times$4 k-point grid for pure $\alpha$-boron. Both rhombohedral and triclinic lattices were examined to find the optimal structure. Hydrogen in boron was modelled using a 2$\times$2$\times$2 k-point grid for a triclinic cell containing 8 B$_{12}$ icosahedra, {\it i.e.} B$_{96}$H$_n$, $n=1,2$.
H$^+$ and H$^-$ were modelled in charge neutral unit cells by the inclusion of a uniform countering background charge. Diffusion barriers were determined using the nudged elastic band method \cite{Henkelman2000}. Saddle points were checked via additional energy double derivative calculations giving a single negative frequency.
The majority of the energies considered below are relative, for example comparison of different stable sites for H$^+$ within the lattice, or calculated migration barriers.
There are a number of different ways to correct absolute binding energies to take into account interaction between charged defects in neighbouring unit cells, and the literature is far from being clear on the best approach. While the standard reference for some time has been \citet{Makov1995}, more recently other approaches are under development\cite{Freysoldt2009,Lany2009}.
The base interaction energy can be estimated from the Madelung energy for an array of point charges with a neutralizing background as $E = - \alpha q^2 / (2 \epsilon L)$, and a reasonable approximation to the charged cell correction is then given in \cite{Lany2009} as $\approx 2 E / 3$.
While corrections can be reasonably large for defects with higher charge states, as our H ions are charge q=$\pm$1 and E varies with q$^2$ these corrections are small. Our repeated unit cell sizes used for the H insertion calculations are also large giving
$L = \Omega^{1/3} = 8.74\AA$. Taking the $\beta$-boron dielectric constant $\epsilon \approx$10 and an overestimate for the Madelung constant $\alpha$ of 10 only results in corrections to formation energies of $\approx$0.04eV. This correction is much smaller than any of the energy differences we calculated between different hydrogen arrangements.
\section{Hydrogen addition to isolated B$_{12}$ icosahedron}
\begin{figure}
\includegraphics[width=8cm]{fig1.eps}
\caption{Diffusion barrier (eV) for H to leave an isolated B$_{12}$ cluster. \label{neb}}
\end{figure}
As a basic element of the $\alpha$-boron structure we first examine an isolated B$_{12}$ cluster. Without hydrogen it forms a perfect icosahedron with B-B bondlengths of $d_{\mathrm{B}_{12}}=1.68$ \AA, in agreement with literature values (1.7 {\AA}\cite{Caputo2009}, 1.71 {\AA}\cite{Fujimori1997}). Adding hydrogen to the B$_{12}$, the energetically most stable position is with hydrogen bound to the outside of B$_{12}$ ($d_{\mathrm{B-H}}=1.19$ {\AA}), deforming the icosahedron giving bondlengths in the range $d_{\mathrm{B}_{12}}=1.62-1.74$ {\AA}. An earlier study found hydrogen stable in the center of the B$_{12}$ cluster \cite{Hayami1999}, however our energy calculations (Figure \ref{neb}) show that this site is actually a metastable maximum (with a slightly expanded B$_{12}$ to $d_{\mathrm{B}_{12}}=1.72-1.73$ {\AA}). Any off-site motion moves the H to an interior site near a B-B-B triangle of the B$_{12}$ ($d_{\mathrm{B}_{12}}=1.69-1.83$ {\AA}). The energy difference between these two sites is only 0.17 eV, while the position outside the B$_{12}$ cluster is 6.05~eV more stable than the center position. Hydrogen inside the B$_{12}$ cluster migrating to the most stable configuration only has to overcome a barrier energy of $\Delta E_B = 0.0042$~eV. This small value indicates that hydrogen atoms will sit covalently bonded to the outside of isolated B$_{12}$ clusters.
\section{Single hydrogen atoms in $\alpha$-boron}
We next consider $\alpha$-boron. For the following discussion we follow the nomenclature to describe inter-cage bonding adopted by \cite{Vast1997,Oganov2009,Fujimori1999}, {\it i.e.} inter-layer bonds are referred to as 2-center (2c) bonds, intra-layer bonds as 3-center (3c) bonds (see Figure \ref{Hinalphapic}). It should not be confused with the nomenclature developed for boron-hydrogen molecular compounds by Lipscomb {\it et al} \cite{Dixon1977,Brown1977,LonguetHiggins1949}, notably their chemical two- and three-center bonds refer to the number of boron atoms involved in chemical bonding, rather than the topology of the boron cage lattice as in our case.
Using a rhombohedral supercell we calculated lattice parameters to be $\alpha= 58.15^\circ$ and $a= 4.98$ {\AA} and B-B bond lengths for the B$_{12}$ icosahedra of $d_{\mathrm{B}_{12}}=1.72-1.78$ {\AA} with inter-icosahedral two-center bonds (2c) $d_{2c}=1.65$ {\AA} and three-center bonds (3c) $d_{3c}=1.98$ {\AA}. The resulting optimised structure is in very good agreement with calculations from \citet{Vast1997} and very similar to other cases \cite{Kawai1991,Masago2006}. Bond lengths using a very similar triclinic supercell ($a=4.83$ {\AA}, $b=4.83$ {\AA}, $c= 4.98$ {\AA}, $\alpha= 60.12^\circ$, $\beta=60.95^\circ$, $\gamma=60.97^\circ$), also with a $4 \times 4 \times 4$ k-point mesh found exactly the same inter-icosahedral 2c- and 3c-bonds lengths and $d_{\mathrm{B}_{12}}=1.72-1.77$ {\AA} for the B$_{12}$ icosahedra. The volume of the triclinic cell is 0.09 \% smaller than the rhombohedral, and the total energy of the cell is 0.0106~eV lower. We therefore used the triclinic cell for all subsequent implantation calculations.
\begin{figure}[b]
\begin{flushleft}
(a)
\end{flushleft}
\vspace{-0.2cm}
\includegraphics[width=7cm]{fig2.eps}\\
\vspace{-4.5cm}
\begin{flushleft}
(b)
\end{flushleft}
\vspace{3.5cm}
\caption{(a) stable sites B-D of H$^0$ (shown white) in the $\alpha$-boron structure with three-center bonds (3c), (b) site A near a two-center bond (2c) shown in a sideview of the $\alpha$-boron structure. \label{Hinalphapic}}
\end{figure}
A single neutral H atom was then added at different sites and the structures optimised again, this time with fixed lattice parameters.
For H$^0$ four stable sites (A-D) were found, as shown in Figure \ref{Hinalphapic}a,b. In site A the H atom is located near a 2c-bond with bond lengths $d_{\mathrm{H_A-B}}=1.35$ {\AA} and $d_{\mathrm{H_A-B}}=1.26$ {\AA}. In site B the H atom sits centered over a boron triangle of 3c-bonds between the icosahedra ($d_{\mathrm{H_B-B}}=1.27$ {\AA}). Site C is near site B, but the H atom lies directly over a 3c-bond between two icosahedra ($d_{\mathrm{H_C-B}}=1.41$ {\AA} for both).
As for isolated B$_{12}$ we also found a much less stable metastable site within the B$_{12}$ cage (see Table \ref{Hinalpha}), position D ($d_{\mathrm{H_D-B}}=1.28$ {\AA} for all three), where the hydrogen lies near a boron triangle pointing towards a 2c-bond. When placed in other sites within the B$_{12}$ icosahedron (including the center site) the hydrogen moves to more stable locations outside the icosahedron without any energy barrier. The energy barriers to move between site D and either site A or B (Table \ref{barr}) are very small (0.13~eV and 0.09~eV respectively), and thus as for the isolated B$_{12}$, H will not remain within the cage except possibly at liquid helium temperatures. This is in agreement with recent calculations for much higher hydrogen densities (1 per B$_{12}$) \cite{Caputo2009}, although we find a different final stable location outside the cage.
As well as the space at icosahedra centers the $\alpha$-boron lattice also has an interstitial void space between the B$_{12}$ layers (see Figure \ref{Hinalphapic}), however once again hydrogen is expected to be metastable here and will displace to more stable neighbouring bond sites.
\begin{table}
\caption{Sites for H$^0$, H$^+$ and H$^-$ in $\alpha$-boron and their calculated stability $\Delta$E (eV) relative to the most stable site at that charge state (as shown in Fig. \ref{Hinalphapic}). \label{Hinalpha}}
\begin{ruledtabular}
\begin{tabular}{l|c c c}
Site & H$^0$ $\Delta$E (eV) & H$^+$ $\Delta$E (eV)& H$^-$ $\Delta$E (eV)\\
\hline
$A$/$A_+$/$A_-$& 0.0 & 0.0 & +0.95 \\
$B$/$B_+$/$B_-$ & +0.17 & +0.48 & 0.0\\
$C$/-/- & +0.24 & - & - \\
$D$/$D_+$/$D_-$ & +2.35 & +2.31 & +3.39\\
-/-/$B*$ & - & - & +0.56 \\
Inter. Void & +0.60 & +1.82 & +0.73 \\
\end{tabular}
- : site is not stable in this configuration
\end{ruledtabular}
\end{table}
\begin{figure}
(a) \hspace{4.5cm} (b)\\
\includegraphics[width=8cm]{fig3.eps}
\caption{The most stable sites for (a) H$^-$ (site $B_-$) and (b) H$^0$ (site $B$) compared with site $B*$ for H$^-$ over and under a 3c bond triangle. \label{H-inalphapic}}
\end{figure}
We next determine the stable sites for H$^+$ and H$^-$ (sites marked as $A_x$, $B_x$,.. with $x=+,-$). For H$^+$ stable sites were found to be comparable to that of neutral hydrogen, with the most stable location near a 2c-bond (site $A_+$, $d_{\mathrm{H_{A+}-B}}=1.27$ {\AA} for both). Site $B_+$ is also stable ($d_{\mathrm{H_{B+}-B}}=1.46$ {\AA} for all three bonds, directly centered between the 3c bond triangle B atoms and inducing an enlargement of the 3c-bonds underneath to $d_{3c}=2.26$ {\AA}), but with a relative energy difference of +0.48~eV to site $A_+$. For H$^+$ site C is not stable (see Table \ref{Hinalpha}).
H$^-$ behaves differently and this time the most stable site is $B_-$ (see Table \ref{Hinalpha}). As for H$^+$ the hydrogen sits between the icosahedra and induces a separation of the icosahedra with $d_{3c}=2.28$ {\AA} (see Figure \ref{H-inalphapic}a). However H$^-$ forms a shorter bond ($d_{\mathrm{H_{B-}-B}}=1.2$ {\AA}) and sits off-centered over the triangle space, pointing more into the interstitial void.
Unlike H$^0$ and H$^+$ there is also a new site called $B*$ near the B$_{12}$ icosahedron on the other side of the triangle. On this side of the triangle the H$^-$ lies over a boron bond of the B$_{12}$-icosahedron with bond lengths $d_{\mathrm{H_{B*}-B}}=1.35$ {\AA} and $d_{\mathrm{H_{B*}-B}}=1.50$ {\AA} (the difference between site $B*$ and site B for H$^0$ is shown in Figure \ref{H-inalphapic}b). Site $A_-$ is a metastable position switching ether to $B_-$ or $B*$. Site C was not found to be stable for H$^-$.
As for H$^0$, site $D_+$ and $D_-$ (the off-centered site within the icosahedron) are energetically unfavourable for both H$^+$ and H$^-$, and we suppose hydrogen leaves the cage to sit in position $A_+$ or respectively $B_-$.
Thus in summary, the most stable site for H$^+$ and H$^0$ in $\alpha$-boron is just off the 2c-bond between layers of B$_{12}$ icosahedra, while H$^-$ prefers to sits in-plane between three icosahedra, slightly separating them.\\
\begin{table
\caption{Calculated diffusion barriers $\Delta E_B$ (eV) for H$^0$, H$^+$ and H$^-$ in $\alpha$-boron. (Path indicates the barrier which has to overcome passing from site $x$ to site $y$: $x \rightarrow y$) \label{barr}}
\begin{ruledtabular}
\begin{tabular}{c c | c c | c c}
\multicolumn{2}{c|}{H$^0$} & \multicolumn{2}{c|}{H$^+$} & \multicolumn{2}{c}{H$^-$} \\
\hline
Path & $\Delta E_B$& Path & $\Delta E_B$& Path & $\Delta E_B$\\
$D\rightarrow A$& 0.13 &$B_+ \rightarrow A_+$& 0.44 &$A_- \rightarrow B_-$ & 0.03\\
$D\rightarrow B$& 0.09 & & &$B* \rightarrow B_-$& 0.39 \\
$B\rightarrow A$ & 0.08 & & & &\\
\end{tabular}
\end{ruledtabular}
\end{table}
We next calculate diffusion barriers between the most stable sites for isolated H$^0$, H$^+$ and H$^-$ (see Table \ref{barr}).
H$^0$ has a very low diffusion barrier of 0.25~eV ($0.17+0.08$) and hence will be highly mobile in the inter-layer region of $\alpha$-boron. H$^+$ also migrates between interlayer sites but with a diffusion barrier of 0.92~eV ($0.48+0.44$). H$^-$ however migrates within the layers between 3c-bond sites, also with a barrier of 0.95~eV ($0.56+0.39$).
\section{Hydrogen pairs and H$_2$}
Comparison of the total energies for two isolated H atoms in boron shows that the most stable configuration is not with two H$^0$ but rather one H$^+$ and one H$^-$, {\it i.e.} H in $\alpha$-boron is a {\it negative U} center (see Table \ref{tab:H2}). This suggests that low density atomic H will adopt an overall charge neutral mixture of H$^+$ and H$^-$ centers, with no net doping of the lattice. While neutral H would be highly mobile within the lattice, the thermodynamically preferred coexistence of H$^+$ and H$^-$ centers should only just be mobile at room temperature. We note that, depending on the defect density and associated degree of electronic coupling between them via the host boron lattice, it may be possible to photoexcite H$^+$/H$^-$ into a metastable H$^0$/H$^0$ state, as has been shown recently for N$_{2}$H centers in diamond \cite{Goss2010}.
While doping $\alpha$-boron could encourage charge compensation by interstitial H and result in a concentration bias of either H$+$ or H$^-$, in preliminary test calculations we found neither potassium or vanadium were stable within the B$_{12}$ icosahedra.
H$^+$/H$^-$ centers will experience a Coulombic attraction and we might expect H$^+$ and H$^-$ to migrate towards one another. Table~\ref{tab:H2} shows that placing two H atoms in the same inter-icosahedral triangle (not dissimilar to H$_2^*$ in silicon \cite{Jackson1991}) lowers the net system energy by 0.84~eV compared to an infinitely separated H$^+$/H$^-$ pair. These neighbouring pairs can then combine to create an H$_2$ molecule, releasing a further 0.79~eV, which sits in the void space between icosahedral layers (see Figure \ref{H2pic}). However it is highly constrained, and the system could lower its energy by a further 1.70~eV if H$_2$ was able to migrate to the boron surface and escape.
The $\alpha$-boron interlayer space contains two types of interlayer void. The first and largest lies between 3c bond triangle sites in neighbouring layers, where the H$_2$ sits (see Figure \ref{H2pic}).
These alternate with a smaller space between a 3c bond triangle in one layer and the base of an icosahedron in the other. Any interlayer diffusion path must necessarily pass through both sites.
Our calculations show that H$_2$ is unable to migrate as a molecule, instead dissociating into a H$^+$/H$^-$ pair which then diffuse through this smaller void before recombining. The NEB calculations give a total barrier of $\approx$2.2~eV, consisting of an $\approx$1.4~eV dissociation energy followed by $\approx$0.9~eV migration barriers for each hydrogen. This means H$_2$ will be trapped and immobile in the large voids until high temperatures. We note that we did not find any other stable sites for H$_2$ in the crystal, suggesting a maximum possible H$_2$ concentration in $\alpha$-boron of B$_{24}$H$_2$.
The initial H$_2$ separation into H$^+$/H$^-$ is barrierless, showing that H$_2$ formation from H ions on neighbouring sites can also occur spontaneously.
We note that we know of no other material where H$_2$ is stable yet migrates by dissociation. The behaviour lies between that of diamond, where the H$_2$ molecule is not stable, and silicon, where H$_2$ is stable and can migrate as a molecule.
\begin{table}
\caption{\label{tab:H2}Comparison of stability for different arrangements of hydrogen
pairs within $\alpha$-boron. H$^x$ refers to hydrogen in its most stable location
with this charge state ({\it i.e.} inter-cage sites).}
\begin{ruledtabular}
\begin{tabular}{l c}
Hydrogen Arrangement & $\Delta E$ (eV)\\
\hline
H$_2$ molecule separated from boron & -\\
H$_2$ inside the boron lattice & +1.70\\
H$^+$/H$^-$ ``intimate pair'' (same triangle) & +2.49\\
H$^+$/H$^-$ infinitely separated in the lattice & +3.33\\
H$^0$/H$^0$ infinitely separated in the lattice & +4.12\\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{figure}
(a) \hspace{4.5cm} (b)\\
\includegraphics[width=8.5cm]{fig4.eps}
\caption{$H_2$ in the hexagonal interstitial site between icosahedral layers, (a) side and (b) top view. \label{H2pic}}
\end{figure}
\section{Discussion and Conclusions}
Our results for the $\alpha$-phase allow us to speculate on hydrogen behaviour in the more
complex $\beta$-boron phase. $\beta$-boron has no equivalents of the $\alpha$-boron 3c bond triangles,
however bonds with similar lengths to the 2c- and 3c- bonds exist. Thus we would
expect no equivalent of the B, B$^*$ and B$^-$ sites for hydrogen, and the most stable sites are likely instead
to be similar to A (2c-bonds) or C (3c-bonds). This agrees with results from implantation of positive muonium into
$\beta$-boron \cite{Cox2009}. We can also suppose that H will
not sit within the B$_{12}$ icosahedra in the $\beta$-boron phase. Unlike $\alpha$-boron there are also some
under-coordinated partial vacancy sites likely to trap hydrogen, and given the larger connected void
spaces we assume H$_2$ migration will have a significantly lower barrier.\\
Table \ref{tab:H2} shows that hydrogen within the $\alpha-$boron lattice is less stable than isolated gas phase H$_2$.
Thus hydrogen incorporation within the lattice will only occur through kinetic restrictions, {\i.e.} if hydrogen becomes trapped in the lattice,
either during growth or through implanation, and is then not able to escape. Since $\beta-$boron has larger void spaces we assume that H$_2$
will be more stable in the $\beta-$boron void sites. Thus the presence of significant quantities of hydrogen within the $\alpha-$boron lattice
may serve to destabilise it with respect to the less dense $\beta-$boron phase.
This may be a contributing factor to the observed difficulty in synthesising $\alpha-$boron, as well as the $\alpha\rightarrow\beta-$boron phase transition
(we note that molecular hydrogen in other crystalline semiconductors such as silicon is often difficult to identify).
Thus we have now a first picture of hydrogen behaviour in pristine $\alpha$-boron.
At low densities H atoms will form a reservoir of H$^+$ and H$^-$ ions in the inter-icosahedral space, resulting in a charge neutral crystal.
Once these become mobile, either through thermal activation or via excitation into highly mobile metastable H$^0$, they will diffuse together, creating interstitial molecular H$_2$. This will remain trapped in interstitial void spaces until at very high temperatures when it will diffuse out. These results suggests that hydrogen doping of boron will be difficult and probably only possible through ion implantation rather than gas phase impregnation and thus $\alpha$-boron is unlikely to serve as a useful hydrogen storage material.
There may also be the possibility of H diffusion and resultant H$_2$ formation at very low temperatures due to quantum tunnelling.
\begin{acknowledgments}
PhW and CPE thank the NANOSIM\_ GRAPHENE project n$^{\circ}$ANR-09-NANO-016-01 funded by the French National Agency (ANR) in the frame of its 2009 programme in Nanosciences, Nanotechnologies \& Nanosystems (P3N2009). We thank the CCIPL where some of these calculations were performed.
\end{acknowledgments}
\bibliographystyle{apsrev4-1}
|
\section{Introduction}
This article is devoted to the study of the representation theory of super Yang-Mills algebras.
Let us briefly recall the definition of the super Yang-Mills algebras.
Given two nonnegative integers $n, s \in \NN_{0}^{2} \setminus \{ (0,0) \}$, and a collection of $(s \times s)$-matrices $\Gamma_{a,b}^{i}$,
for $i = 1, \dots, n$ ($a,b = 1, \dots, s$),
the \emph{super Yang-Mills algebra} $\ym(n,s)^{\Gamma}$ over an algebraically closed field $k$ of characteristic zero is defined as
the quotient of the free super Lie algebra $\f(x_{1},\dots,x_{n},z_{1},\dots,z_{s})$, for even indeterminates $x_{1}, \dots, x_{n}$ and odd ones
$z_{1}, \dots, z_{s}$, by the (homogeneous) relations given by
\begin{align*}
r_{0,i} &= \sum_{j=1}^{n} [x_{j},[x_{j},x_{i}]] - \frac{1}{2} \sum_{a,b = 1}^{s} \Gamma^{i}_{a,b} [z_{a},z_{b}],
\\
r_{1,a} &= \sum_{i=1}^{n} \sum_{b = 1}^{s} \Gamma_{a,b}^{i} [x_{i},z_{b}],
\end{align*}
for $i = 1, \dots, n$ and $a = 1, \dots, s$, respectively.
It can also be regarded as a graded Lie algebra with $\deg(x_{i})=2$, for $i = 1, \dots, n$, and $\deg(z_{a})=3$, for $a = 1, \dots, s$.
The associative enveloping algebra $\U(\ym(n,s)^{\Gamma})$ will be denoted $\YM(n,s)^{\Gamma}$.
They have been previously considered by M. Movshev and A. Schwarz in \cite{Mov05} and \cite{MS06}.
Also, the case $s = 0$ (and $n \geq 2$) leads to the definition of Yang-Mills algebra given by A. Connes and M. Dubois-Violette in \cite{CD02}.
Omitting the trivial cases with $n=0$, or the already known ones with $s=0$,
we shall see that, for each $(n,s) \in \NN^{2}$, it is noetherian if and only if $n = 1$ and $s = 1, 2$.
However, it is coherent for all values of the parameters $(n,s)$ (see Remark \ref{rem:coh}).
From the physical point of view, it can be seen that the components of the covariant derivative and the dual spinor field of a supersymmetric gauge theory
on the Minkowski space provide a representation of the corresponding super Yang-Mills algebra (\textit{cf.} \cite{DF99s}, and see also Remark \ref{rem:phys}).
Otherwise stated, these super algebras yield a ``background independent'' formulation of the supersymmetric gauge theory in physics.
They can also be regarded in order to provide a noncommutative version of it.
Our interest in them comes then in order to shed more light in this direction.
The main result of this article may be formulated as follows:
\begin{theorem*}
Let $n, s, p, q \in \NN$ be positive integers, satisfying $n \geq 3$.
We suppose further that either $p \geq 3$, or $p = 2$ and $q \geq 2$.
Then, there exists a surjective homomorphism of super algebras
\[ \U(\ym(n,s)^{\Gamma}) \twoheadrightarrow \Cliff_{q}(k) \otimes A_{p}(k), \]
where $\Cliff_{q}(k) \otimes A_{p}(k)$ denotes the Clifford-Weyl super algebra.
Furthermore, there exists $l \in \NN$ such that we can choose this morphism in such a way that it factors through the quotient
$\U(\ym(n,s)^{\Gamma}/F^{l}(\ym(n,s)^{\Gamma}))$, where $F^{l}(\ym(n,s)^{\Gamma})$ is the Lie ideal of $\ym(n,s)^{\Gamma}$ formed by the elements
of degree greater than or equal to $l+2$.
\end{theorem*}
We would like to remark that the Clifford-Weyl super algebra $\Cliff_{q}(k) \otimes A_{p}(k)$ appearing in the previous theorem has the $\ZZ/2\ZZ$-grading given by usual grading of the Clifford (super) algebra
$\Cliff_{q}(k)$ and by considering the Weyl algebra $A_{p}(k)$ to be concentrated in degree zero (see \cite{Her10}, Example 1.2).
This differs from the grading of the ``Clifford-Weyl algebras'' $\mathcal{C}(q, 2 p)$ considered in \cite{MPU09}, since in that case the Weyl algebra has also a nontrivial homogeneous component of odd degree.
In order to prove the theorem we needed to extend the so called \emph{Dixmier map} for nilpotent Lie algebras to the case of nilpotent super Lie algebras,
which was done in \cite{Her10}.
Even tough this article may be considered as an extension and generalization of the results proved in \cite{HS10},
it also deals with several difficulties and differences with the latter, which mainly follow from the fact that the study of the
enveloping algebras of super Lie algebra has various differences with the case of enveloping algebras of Lie algebras.
To mention just a few, the super Yang-Mills algebras are not Koszul, at least not for any definition we are aware of,
even though they behave quite the same,
there are not any a priori morphism between different super Yang-Mills algebras for arbitrary $\Gamma$'s (see the paragraph at the end of Subsection \ref{subsec:defi}),
there are several differences between the Dixmier map of nilpotent Lie algebras and of nilpotent super Lie algebras
(\textit{e.g.} the super dimension of a polarization at an even functional does not determine the weight of the ideal that it defines, which explains the phenomena that appears in Remark \ref{rem:otrosn}, etc),
the enveloping algebra of a super Lie algebra is not necessarily semiprimitive, and the determination of its radical is usually a highly nontrivial task, etc.
\bigskip
The contents of the article are as follows.
In Section \ref{sec:generalities} we recall the definition and several elementary properties of super Yang-Mills algebras,
some of them with a physical flavour.
In particular, we show in Subsection \ref{subsec:pot} that the typical odd supersymmetries of the supersymmetric gauge theories considered in physics also appear, under the same assumptions as there,
as super derivations of the super Yang-Mills algebra.
In Section \ref{sec:hom} we study the homological properties of this family of super algebras.
In fact, Subsection \ref{subsec:projres} provides a complete description of the minimal projective resolution of the trivial module $k$
over the graded Lie algebra $\ym(n,s)^{\Gamma}$, for $(n,s) \neq (1,0),(1,1)$,
which satisfy a property similar to koszulity.
Using a procedure similar to the one employed by R. Berger and N. Marconnet in \cite{BM06}, Section 4, we furthermore obtain
the minimal projective resolution of the super Yang-Mills algebra $\YM(n,s)^{\Gamma}$, considered as a graded algebra, in the category of bimodules,
for the same set of indices.
From the particular description of these minimal projective resolutions we prove that $\YM(n,s)^{\Gamma}$ AS-regular in the sense of \cite{MM11}
and graded Calabi-Yau, for $(n,s) \neq (1,0), (1,1)$.
We later derive some consequences, computing in particular the Hilbert series of both $\ym(n,s)^{\Gamma}$ and $\YM(n,s)^{\Gamma}$.
Moreover, we prove that, for $n \geq 2$, or $n = 1$ and $s \geq 3$, the super Yang-Mills algebra $\ym(n,s)^{\Gamma}$ contains a finite codimensional Lie ideal
which is a free super Lie algebra.
This is done using simpler methods than the ones used in \cite{HS10}, Section 3.
From this fact, we derive several consequences, and in particular that these graded algebras are not noetherian, but they are coherent.
Finally, in section \ref{sec:main} we prove our main result, Theorem \ref{teo:yangmillsweyl},
and describe the families of representations appearing in this way.
We would like to thank A. Solotar for several suggestions and remarks.
\section{Generalities}
\label{sec:generalities}
In this first section we fix notations and recall some elementary properties of what we call the super Yang-Mills algebras.
\subsection{Definition}
\label{subsec:defi}
Throughout this article $k$ will denote an algebraically closed field of characteristic zero.
The main convention and notations on super vector spaces, super algebras and modules over them (and also for the graded analogous ones)
that we follow are the same as in \cite{Her10}, to which we refer.
Unless otherwise stated, a module over an algebra (resp. a graded algebra, a super algebra), will always denote a left module.
Moreover, we consider the category of modules over graded (resp. super) algebras provided with homogeneous linear morphisms of degree zero.
It is also endowed with the shift functor $(\place)[1]$, defined by $(M[1])_{n} = M_{n+1}$, for $n \in \ZZ$ in the graded case, and
$n \in \ZZ/2\ZZ$ in the super case.
Given two modules $M$ and $N$ over a graded (resp. super) algebra $A$, $\mathrm{hom}_{A}(M,N)$ will stand for space of morphisms
in the previously described categories.
Furthermore, the internal space of morphisms is given by $\mathcal{H}om_{A}(M,N) = \oplus_{i \in G} \mathrm{hom}_{A}(M,N[i])$, where $G = \ZZ$ or $G = \ZZ/2\ZZ$ for the graded or the super case, respectively.
We fix the following setup.
Let $V = V_{0} \oplus V_{1}$ be a super vector space over $k$ of super dimension $(n,s) \in \NN_{0}^{2}$ with $n + s > 0$,
such that the even part $V_{0}$ is provided with a nondegenerate symmetric bilinear form $g$.
Note that the algebraic group $\SO(V_{0},g)$, and hence the Lie algebra $\so(V_{0},g)$, acts on $V_{0}$ (with the standard action).
We shall also write $V = V(n,s)$, $V_{0}=V(n)_{0}$ or $V_{1}=V(s)_{1}$ if we want to stress the (super) dimension.
We suppose further that there exists a map of the form $\Gamma : S^{2}V^{*}_{1} \rightarrow V_{0}$.
Choose a (homogeneous) basis $\B = \B_{0} \cup \B_{1}$ of $V$, where $\B_{0} = \{ x_{1}, \dots, x_{n} \}$ and $\B_{1} =\{ z_{1}, \dots, z_{s} \}$,
with $|x_{i}| = 0$, for all $i = 1, \dots, n$, and $|z_{a}| = 1$, for all $a = 1, \dots, s$,
and let $\B^{*} = \B^{*}_{0} \cup \B^{*}_{1}$, where $\B^{*}_{0} = \{ x_{1}^{*}, \dots, x_{n}^{*}\}$ and $\B^{*}_{1} = \{z_{1}^{*}, \dots, z_{s}^{*} \}$,
be the dual basis of $V^{*}$.
Set $\Gamma_{a,b}^{i} = x_{i}^{*}(\Gamma(z_{a}^{*},z_{b}^{*}))$, for $i = 1, \dots, n$ and $a, b = 1, \dots, s$, and
$g^{-1}$ the \emph{inverse} nondegenerate symmetric bilinear form on $V^{*}_{0}$,
\textit{i.e.} $g^{-1}$ is the bilinear form on $V^{*}_{0}$ defined as the image of $g$ under
the $k$-linear isomorphism $V_{0} \rightarrow V^{*}_{0}$ given by $v \mapsto g(v,\place)$.
It is easy to see that the matrix of $g^{-1}$ with respect to the dual basis $\B^{*}_{0}$ is the inverse of the matrix of $g$ with respect to the basis $\B_{0}$,
and we further write $g^{i,j} = g^{-1}(x_{i}^{*},x_{j}^{*})$ and $g_{i,j} = g(x_{i},x_{j})$.
If $\f(V)$ denotes the free super Lie algebra generated by the super vector space $V$,
the \emph{super Yang-Mills algebra} is defined as the quotient
\[ \ym(V,g)^{\Gamma} = \f(V)/\cl{R(V,g)^{\Gamma}}, \]
where $R(V,g)^{\Gamma}$ is the super vector space inside $\f(V)$ spanned by the elements
\begin{equation}
\label{eq:rel}
\begin{split}
r_{0,i} &= \sum_{j,l,m=1}^{n} g^{j,l} g^{i,m} [x_{j},[x_{l},x_{m}]] - \frac{1}{2} \sum_{a,b = 1}^{s} \Gamma^{i}_{a,b} [z_{a},z_{b}],
\\
r_{1,a} &= \sum_{i=1}^{n} \sum_{b = 1}^{s} \Gamma_{a,b}^{i} [x_{i},z_{b}],
\end{split}
\end{equation}
for $i = 1, \dots, n$ and $a = 1, \dots, s$, respectively.
We consider the universal enveloping algebra $\YM(V,g)^{\Gamma} = \U(\ym(V,g)^{\Gamma})$ of $\ym(V,g)^{\Gamma}$, and also call it the
\emph{(associative) super Yang-Mills algebra}.
By definition, it is the super algebra given by the quotient of the tensor algebra $TV(n,s)$ of the super vector space $V(n,s)$ by the two-sided ideal generated by
the same super vector subspace $R(V(n,s),g)^{\Gamma}$, now seen inside of $F^{3}TV(n,s)$,
where $\{ F^{\bullet}TV(n,s) \}_{\bullet \in \NN_{0}}$ denotes the canonical (increasing) filtration of the tensor algebra.
Note that $\U(\ym(n,s)^{\Gamma})$, being the enveloping algebra of a super Lie algebra, need not be a domain, whereas
$\U(\ym(n,s)^{\Gamma}_{0})$ is always so.
However, we shall see below that $\U(\ym(n,s)^{\Gamma})$ is also a domain.
It is direct to check that $R(V,g)^{\Gamma}$ is independent of the choice of the homogeneous basis $\B$, so we may suppose that $\B_{0}$ is orthonormal,
in which case the relations simplify to give
\begin{equation}
\label{eq:relort}
\begin{split}
r_{0,i} &= \sum_{j=1}^{n} [x_{j},[x_{j},x_{i}]] - \frac{1}{2} \sum_{a,b = 1}^{s} \Gamma^{i}_{a,b} [z_{a},z_{b}],
\\
r_{1,a} &= \sum_{i=1}^{n} \sum_{b = 1}^{s} \Gamma_{a,b}^{i} [x_{i},z_{b}],
\end{split}
\end{equation}
which we will assume from now on.
Therefore, if the super vector space $V$ has super dimension $(n,s)$, we may also denote the super Yang-Mills algebras by
$\ym(n,s)^{\Gamma}$ and by $\YM(n,s)^{\Gamma}$, respectively.
We may also write $R(n,s)^{\Gamma}$ instead of $R(V,g)^{\Gamma}$.
Note that they generalize the Yang-Mills algebras defined by A. Connes and M. Dubois-Violette in \cite{CD02}, since $\ym(n) = \ym(n,0)^{0}$,
and have been previously considered by M. Movshev and A. Schwarz in \cite{Mov05} and \cite{MS06}.
We would also like to mention that, as far as we know, there is no direct relation between the previously defined super algebra and the superized versions of the Yang-Mills algebra defined in \cite{CD07}, Section 1.4,
and, more generally, in \cite{HKL08}, Example 3.2.
We are aware of the fact that the superized version of the Yang-Mills algebra in \cite{CD07}, Section 1.4, was also called super Yang-Mills algebra there,
but we decided to use the same name just due to the connection to supersymmetric Yang-Mills theory in physics.
We say that the super Yang-Mills $\ym(n,s)^{\Gamma}$ is \emph{equivariant} if $V_{1}$ is a representation of the Lie algebra $\so(V_{0},g)$
such that the map $\Gamma : S^{2}V^{*}_{1} \rightarrow V_{0}$ is $\so(V_{0},g)$-equivariant, and there also exists an $\so(V_{0},g)$-equivariant map
$\tilde{\Gamma} : S^{2}V_{1} \rightarrow V_{0}$, which satisfy the following condition.
Rewritting $\Gamma$ and $\tilde{\Gamma}$ as elements $\gamma \in \Hom (V_{0},\Hom(V_{1}^{*},V_{1}))$ and $\tilde{\gamma} \in \Hom (V_{0},\Hom(V_{1},V_{1}^{*}))$
defined by $z^{*}_{2}(\gamma(v)(z^{*}_{1})) = g(\Gamma(z^{*}_{1},z^{*}_{2}),v)$, for all $v \in V_{0}$ and $z_{1}^{*}, z_{2}^{*} \in V_{1}^{*}$, and
by $\tilde{\gamma}(v)(z_{1})(z_{2}) = g(\tilde{\Gamma}(z_{1},z_{2}),v)$, for all $v \in V_{0}$ and $z_{1}, z_{2} \in V_{1}$, respectively,
the assumption reads as follows
\begin{equation}
\label{eq:relgammas}
\begin{split}
\tilde{\gamma}(v) \circ \gamma(v) &= g(v,v) \id_{V_{1}^{*}},
\\
\gamma(v) \circ \tilde{\gamma}(v) &= g(v,v) \id_{V_{1}},
\end{split}
\end{equation}
for all $v \in V_{0}$.
In particular, this implies that $\tilde{\Gamma}$ is uniquely determined from $\Gamma$.
If we denote $\tilde{\Gamma}^{i,a,b} = x_{i}^{*}(\Gamma(z_{a},z_{b}))$, for $i = 1, \dots, n$ and $a, b = 1, \dots, s$, we notice that
the conditions \eqref{eq:relgammas} can be rewritten as
\begin{equation}
\label{eq:relgammasind}
\sum_{b = 1}^{s} (\Gamma^{i}_{a,b} \tilde{\Gamma}^{j,b,c} + \Gamma^{j}_{a,b} \tilde{\Gamma}^{i,b,c}) = 2 g^{i,j} \delta_{a,c}.
\end{equation}
We remark the easy fact that an equivariant super Yang-Mills algebra with $s \neq 0$ satisfies \textit{a fortiori} that $s(s+1)/2 \geq n$,
for $V_{0}$ is an irreducible $\so(V_{0},g)$-module and $\Gamma \neq 0$.
Note that if the super Yang-Mills is equivariant, then the action of $\SO(V_{0},g)$ (and so the action of $\so(V_{0},g)$)
on $V$ induces an action by automorphisms of $\SO(V_{0},g)$ (and hence an action by derivations of $\so(V_{0},g)$) both on the tensor super algebra $TV$
and in the free super Lie algebra $\f(V)$, which preserves the corresponding ideal $\cl{R(n,s)^{\Gamma}}$ generated by $R(n,s)^{\Gamma}$ in the tensor algebra
$TV$ and in the free super Lie algebra $\f(V)$, respectively.
As a consequence, we get an action by automorphisms of $\SO(V_{0},g)$, and an action by derivations of $\so(V_{0},g)$,
on both super Yang-Mills algebras $\YM(n,s)^{\Gamma}$ and $\ym(n,s)^{\Gamma}$.
Even though in the examples coming from physics the super Yang-Mills algebra is equivariant, it will be useful to consider a weaker notion.
We say that $\Gamma$, or even the super Yang-Mills algebra, is \emph{nondegenerate}
if $n \neq 0$ and there exists a nonzero linear form $\lambda \in V_{0}^{*}$ such that
$\lambda \circ \Gamma : V_{1}^{*} \otimes V_{1}^{*} \rightarrow k$ is nondegenerate, if $s \neq 0$.
We remark that the super Yang-Mills algebra is nondegenerate if $s = 0$.
Also, note that condition \eqref{eq:relgammasind} implies that an equivariant super Yang-Mills algebra with $s \neq 0$ is always nondegenerate (in fact each matrix $\Gamma^{i}$
is invertible).
From now on, we shall assume that the super Yang-Mills algebras we are considering are nondegenerate.
Without loss of generality, we shall furthermore suppose that, if $s \neq 0$, $\lambda = x_{1}^{*}|_{V_{0}}$ and that $\B_{1}^{*}$ is orthonormal with respect to
$x_{1}^{*}|_{V_{0}} \circ \Gamma$.
This hypothesis of nondegeneracy is though not necessary for most of this subsection and the next one, except for some minor indicated cases,
but it will be necessary (and assumed) from Subsection \ref{subsec:anodesc} on.
As mentioned before, the case where $s = 0$ and $n \geq 2$ has been previously studied in \cite{CD02} (see also \cite{HS10}).
If $n=1$ and $s=0$, then the (super) Yang-Mills algebra is just the one-dimensional abelian Lie algebra, which was not studied in the mentioned articles because of its simplicity.
Moreover, if $n = 2$ and $s=0$, $\ym(2)$ is isomorphic to the \emph{Heisenberg Lie algebra} $\h_{1}$, with generators $x,y,z$,
and relations $[x,y] = z$, $[x,z] = [y,z] = 0$, and, for $n \geq 3$, $\ym(n)$ is an infinite dimensional Lie algebra (see \cite{HS10}, Rem. 3.14).
On the other hand, for $n=0$ (so $s > 0$), the super Yang-Mills algebra is just the free super Lie algebra generated by the odd elements $z_{1}, \dots, z_{s}$.
The nondegeneracy assumption on the $\Gamma$ also implies that $\ym(1,s) \simeq k.x_{1} \oplus \f(V(1))/\cl{\sum_{a=1}^{s} [z_{a},z_{a}]}$, and then
$\YM(1,s) \simeq k[x_{1}] \otimes k\cl{z_{1},\dots,z_{s}}/\cl{\sum_{a=1}^{s} z_{a}^{2}}$.
In particular, $\ym(1,1)^{\Gamma}$ is a supercommutative super Lie algebra of super dimension $(1,1)$,
and $\ym(1,2)^{\Gamma}$ has super dimension $(2,2)$, with basis $x_{1}$, $z_{1}$, $z_{2}$ and $[z_{2},z_{2}]$, where $[z_{1},z_{1}]=-[z_{2},z_{2}]$, and all other brackets vanish.
We remark that, since the case where $s=0$ is trivial,
we will focus ourselves on indices $(n,s) \in \NN \times \NN_{0}$, unless otherwise stated.
As noted in \cite{MS06}, it can be also useful to consider $\ym(n,s)^{\Gamma}$ as an $\NN$-graded Lie algebra, where the elements $x_{i}$ are of degree $2$,
for all $i = 1, \dots, n$, and the elements $z_{a}$ of degree $3$, for all $a = 1, \dots, s$.
Similarly, taking the grading induced by the previous definitions, the associative super Yang-Mills algebra $\YM(n,s)^{\Gamma}$ can also be regarded as an
$\NN_{0}$-graded algebra.
These gradings for both the Lie and associative versions of the super Yang-Mills algebra are exactly the \emph{special gradings}
(opposed to the \emph{usual} ones) considered in \cite{HS10}, Section 2, and \cite{HS10h}, Section 2.1, when $s=0$.
We remark that the underlying super Lie algebra and super algebra of these latter definitions yield the ones given at the beginning.
Moreover, when the super Yang-Mills algebra $\ym(n,s)^{\Gamma}$ is seen as a graded Lie algebra, we may consider the descending sequence of graded ideals $\{ F^{\bullet}\ym(n,s)^{\Gamma} \}_{\bullet \in \NN_{0}}$,
where $F^{j}\ym(n,s)^{\Gamma}$ is the graded vector subspace of $\ym(n,s)^{\Gamma}$ given by elements of degree greater than or equal to $j + 2$.
It is the quotient under the canonical projection of the descending filtration $F^{\bullet}\f(V(n,s))$ of the graded free Lie algebra
given by elements of degree greater than or equal to $j + 2$.
This also induces a descending sequence of ideals of the underlying super Lie algebra of either the graded free Lie algebra $\f(V(n,s))$
or the graded Lie algebra $\ym(n,s)^{\Gamma}$ considered above.
\begin{remark}
\label{rem:phys}
We remark that this super algebra appears naturally when studying supersymmetric gauge field theories.
We will only recall what we need for our explanation, and we refer to \cite{DF99} and \cite{DF99s} for a complete account on this subject.
Let $\check{M}$ denote an $n$-dimensional \emph{Minkowski space} with real vector space of translations $V$ of dimension $n$
and metric $g$ determining a cone $C$ of time-like vectors in $V$, $S$ a real spinorial representation of dimension $s$ of the spin group $\mathrm{Spin}(V,g)$
and $\Gamma : S^{*} \otimes S^{*} \rightarrow V$ a symmetric morphism of representations of $\mathrm{Spin}(V,g)$, which is \emph{positive definite},
\textit{i.e.} $\Gamma(s^{*},s^{*}) \in \bar{C}$, for all $s^{*} \in S^{*}$, and $\Gamma(s^{*},s^{*}) = 0$ only if $s^{*} = 0$.
Since any complex vector bundle $E$ of rank $m$ over $\check{M}$ is trivial, for $\check{M}$ is contractible,
every connection on such bundle is given by a $\gl_{m}(\CC)$-valued $1$-form $\sum_{i=1}^{n} A_{i} dx^{i}$,
and the corresponding covariant derivative is $\nabla_{i} = \partial_{i} + A_{i}$, for $i = 1, \dots, n$.
We also recall that a \emph{dual spinor field} $\lambda$ with values in the Lie algebra $\gl_{m}(\CC)$ is a morphism from $\check{M}$ to
$S^{*} \otimes \gl_{m}(\CC)$, so it can be decomposed in components $\lambda_{a}$ from $\check{M}$ to $\gl_{m}(\CC)$, for $a = 1, \dots, s$.
We may thus see both set of fields $\nabla_{i}$ and $\lambda_{a}$ as sections of the endomorphism bundle of the complex super vector bundle
$E \otimes_{\RR} \Lambda^{\bullet}_{\RR} S^{*}$ on $\check{M}$, where $S$ is considered to be in degree $1$,
$\nabla_{i}$ is even, for all $i = 1, \dots, n$, and $\lambda^{a}$ is odd, for all $a = 1, \dots, s$.
The associated set of \emph{super Yang-Mills equations} is the supersymmetric extension of the usual Yang-Mills equations
and it is given by
\begin{equation}
\label{eq:ym}
\begin{split}
\sum_{j,l,m=1}^{n} g^{i,j} g^{l,m}[\nabla_{j},[\nabla_{l},\nabla_{m}]] &= \frac{1}{2} \sum_{a,b = 1}^{s} \Gamma^{i}_{a,b} [\lambda_{a},\lambda_{b}],
\\
\sum_{i=1}^{n} \sum_{b = 1}^{s} \Gamma_{a,b}^{i} [\nabla_{i},\lambda_{b}] &= 0,
\end{split}
\end{equation}
for $i = 1, \dots, n$ and $a = 1, \dots, s$, respectively (\textit{cf.} \cite{DF99s}, (6.23)).
We remark that the previous identities are considered in the super Lie algebra of endomorphisms of sections of the complex super vector bundle
$E \otimes_{\RR} \Lambda^{\bullet}_{\RR} S^{*}$, so they give a representation of the equivariant super Yang-Mills algebra $\ym(n,s)^{\Gamma}$.
\end{remark}
The Lie ideal $\tym(n,s)^{\Gamma} = F^{1}\ym(n,s)^{\Gamma} = [\ym(n,s)^{\Gamma},\ym(n,s)^{\Gamma}] + \ym(n,s)^{\Gamma}_{1}$
of the super Lie algebra $\ym(n,s)^{\Gamma}$ will be also important in the sequel.
It is obvious to see that it is also a graded Lie ideal of $\ym(n,s)^{\Gamma}$, when considered as a graded Lie algebra.
Notice that $\tym(n,0)^{0}$ coincides with the ideal $\tym(n)$ of $\ym(n) = \ym(n,0)^{0}$ considered in \cite{HS10}.
We remark that $\ym(n,s)^{\Gamma}/\tym(n,s)^{\Gamma} \simeq V(n)_{0}$ is the abelian (super) Lie algebra of super dimension $(n,0)$.
Moreover, we shall deal with the universal enveloping algebra of the Lie ideal $\tym(n,s)^{\Gamma}$, which will be denoted $\TYM(n,s)^{\Gamma}$,
which can be regarded either as a super algebra or as a graded algebra.
We shall also consider the bigger Lie ideal $\hat{\tym}(n,s)^{\Gamma} = \tym(n,s)^{\Gamma} \oplus \bigoplus_{i=3}^{n} k.x_{i}$,
which satisfies that $\ym(n,s)^{\Gamma}/\hat{\tym}(n,s)^{\Gamma} \simeq V(2)_{0}$ is supercommutative, and its enveloping algebra
$\hat{\TYM}(n,s)^{\Gamma} = \U(\hat{\tym}(n,s)^{\Gamma})$.
Occasionally, we will omit the indices $(n,s)$ and $\Gamma$ for the previously defined spaces (and also for the defined below)
in order to simplify the notation if it is clear from the context.
We would like to make a further comment on the relationship between different super Yang-Mills algebras.
One first notes that the canonical projection $V(n,s) \rightarrow V(n)_{0}$ of graded vector spaces induces a surjective morphism of graded algebras
$TV(n,s) \rightarrow TV(n)_{0}$, which sends the even super Yang-Mills relations $r_{0,i}$ to the usual Yang-Mills relations $r_{i}$ described in \cite{HS10},
and the odd super Yang-Mills relations $r_{1,a}$ to zero.
Hence, we obtain a surjective morphism of graded algebras $\YM(n,s)^{\Gamma} \rightarrow \YM(n)$,
where we recall that the Yang-Mills algebras are provided with the special grading.
In an analogous manner, we have a surjective morphism of graded Lie algebras $\ym(n,s)^{\Gamma} \rightarrow \ym(n)$, which obviously maps
$\tym(n,s)^{\Gamma}$ onto $\tym(n)$.
We would like to point out however, that we do not know of any \textit{a priori} given morphism between different super Yang-Mills algebras with $s \neq 0$
when the corresponding $\Gamma$'s are arbitrary.
\subsection{A superpotential formulation}
\label{subsec:pot}
We remark that relations \eqref{eq:relort} can be obtained from the superpotential element given by the class of the homogeneous element
(resp., of degree $8$ if $\ym(n,s)^{\Gamma}$ is considered as a graded Lie algebra)
\begin{equation}
\label{eq:sup}
W = - \frac{1}{4} \sum_{i,j = 1}^{n} [x_{i},x_{j}][x_{i},x_{j}] + \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b = 1}^{s} \Gamma_{a,b}^{i} z_{a}[x_{i},z_{b}]
\end{equation}
in $HH_{0}(TV) = TV/[TV,TV]$, where we recall that we are considering the super (resp. graded) commutator space $[TV,TV]$ of the tensor algebra, \textit{i.e.}
the super (resp., graded) vector space spanned by the elements $u v - (-1)^{|u||v|} v u$, for homogeneous elements $u, v \in TV$.
In this case, the \emph{cyclic derivative with respect to a generator} $v \in \B$, where $\B$ is a basis of (homogeneous elements of) $V$, is the map from $TV/[TV,TV]$ to $TV$ given by the
$k$-linear extension of the following: for an element $w = v_{1} \dots v_{r}$, where each $v_{i} \in \B$ for all $i = 1, \dots, r$, we consider
\[ \frac{\partial \bar{w}}{\partial v} = \sum_{i : v_{i} = v} (-1)^{(|v_{i}| + \dots + |v_{n}|) (|v_{1}| + \dots + |v_{i-1}|)}
v_{i+1} \dots v_{n} v_{1} \dots v_{i-1}. \]
It can be easily checked that in our case $r_{0,i}$ is the cyclic derivative of $W$ with respect to $x_{i}$, and $r_{1,a}$ is the cyclic derivative of $W$ with respect to $z_{a}$.
We have the following proposition, which we suppose it should be well-known, since it is only the algebraic analogous of the physical folkloric result that states that
the Euler-Lagrange equations associated to an action are invariant under a collection of (super)symmetries if the action is invariant under such (super)symmetries.
\begin{proposition}
\label{prop:simpot}
Let $V$ be a graded vector space $V$, and let $\g$ be a graded Lie algebra acting by homogeneous derivations
on the corresponding tensor algebra $TV$.
Regarding $TV$ as a graded Lie algebra, we immediately note that the action of $\g$ on $TV$ induces an action on (the abelianization) $HH_{0}(TV) = TV/[TV,TV]$.
If a homogeneous element $\bar{W} \in HH_{0}(TV)$ belongs to the invariant space $HH_{0}(TV)^{\g}$, then the two-sided ideal generated by the
cyclic derivatives of $W$ (with respect to any basis of $V$) is preserved by $\g$.
\end{proposition}
\noindent\textbf{Proof.}
It suffices to prove the statement for a unique homogeneous derivation $d$ of $TV$ of degree $|d|$, and call $\bar{d}$ the induced map on $HH_{0}(TV)$.
Let $\B$ be a (homogeneous) basis of $V$.
Fix an element $v$ of the basis $\B$ of $V$, and for each $v' \in \B$,
consider the monomial elements $r_{l,v'}^{m}, s_{l,v'}^{m} \in TV$ defined by $d(v') = \sum_{m \in M_{v'}} r_{l,v'}^{m} v s_{l,v'}^{m}$, for all $l \in L_{v',m}$, and
\[ \frac{\partial (\bar{d}(\bar{v}'))}{\partial v} = \sum_{m \in M_{v'}} \sum_{l \in L_{v',m}} (-1)^{|r_{l,v'}^{m}|(|v|+|s_{l,v'}^{m}|)}
s_{l,v'}^{m} r_{l,v'}^{m} \]
hold, where $M_{v'}$ and $L_{v',m}$ are sets of indices.
Set $\B' = \B \setminus \{ v \}$.
We would like to explain the notation, which could seem cumbersome at first glance.
The element $d(v')$ is a sum of monomials $M_{m}$ in $TV$, indexed by $m \in M_{v'}$, and each of these monomials can be written as $M_{m} = r_{l,v'}^{m} v s_{l,v'}^{m}$, for $l \in L_{v',m}$,
which are all the different ways of writting $M_{m}$ such that the formula for the cyclic derivative can be applied to give
\[ \frac{\partial (\bar{d}(\bar{M_{m}}))}{\partial v} = \sum_{l \in L_{v',m}} (-1)^{|r_{l,v'}^{m}|(|v|+|s_{l,v'}^{m}|)} s_{l,v'}^{m} r_{l,v'}^{m}. \]
We claim that the following identity holds
\begin{equation}
\label{eq:difi}
\frac{\partial (\bar{d}(\bar{W}))}{\partial v} = (-1)^{|d||v|} d\Big(\frac{\partial \bar{W}}{\partial v}\Big)
+ \sum_{v' \in \B} \sum_{m \in M_{v'}} \sum_{l \in L_{v',m}} (-1)^{|r_{l,v'}^{m}|(|W|+|s_{l,v'}^{m}| + |v|-|v'|)} s_{l,v'}^{m}
\frac{\partial \bar{W}}{\partial v'} r_{l,v'}^{m}.
\end{equation}
First, note that it only suffices to prove the previous equality for $W$ a monomial,
since if we write $W = \sum_{j \in J} W_{j}$, where $W_{j}$ are monomials (of the same degree as $W$) in $TV$, the sum of each of the identities \eqref{eq:difi} for
$W_{j}$ gives \eqref{eq:difi} for $W$.
Then, we assume that $W$ is a monomial of the form $\prod_{i = 1}^{N} v_{i}$, for $N \in \NN$, such that
$v_{i} \in \B$.
Define $d_{i,j}$ to be $\sum_{l=i}^{j} |v_{l}|$, for $1 \leq i \leq j \leq N$, and zero else.
On the one hand, we have that
\[ \frac{\partial \bar{W}}{\partial v} = \sum_{i : v_{i} = v} (-1)^{d_{1,i-1}d_{i,N}} (\prod_{j=i+1}^{N} v_{j}) (\prod_{j=1}^{i-1} v_{j}), \]
which yields
\begin{equation}
\label{eq:dpW}
\begin{split}
d\Big(\frac{\partial \bar{W}}{\partial v}\Big)
=& \sum_{i : v_{i} = v} (-1)^{d_{1,i-1}d_{i,N}} \Big(\sum_{p = i+1}^{N} (-1)^{|d|d_{i+1,p-1}}
(\prod_{j=i+1}^{p-1} v_{j}) d(v_{p}) (\prod_{j=p+1}^{N} v_{j}) (\prod_{j=1}^{i-1} v_{j})
\\
&+ \sum_{p = 1}^{i-1} (-1)^{|d|(d_{i+1,N}+d_{1,p-1})}
(\prod_{j=i+1}^{N} v_{j}) (\prod_{j=1}^{p-1} v_{j}) d(v_{p}) (\prod_{j=p+1}^{i-1} v_{j}) \Big).
\end{split}
\end{equation}
On the other hand, from
\begin{align*}
d(W) &= \sum_{p=1}^{N} (-1)^{|d|d_{1,p-1}} (\prod_{j=1}^{p-1} v_{j}) d(v_{p}) (\prod_{j=p+1}^{N} v_{j})
\\
&= \sum_{p=1}^{N} \sum_{m \in M_{v_{p}}} (-1)^{|d|d_{1,p-1}} (\prod_{j=1}^{p-1} v_{j}) r_{l,v_{p}}^{m} v s_{l,v_{p}}^{m} (\prod_{j=p+1}^{N} v_{j}),
\end{align*}
we get that
\begin{align*}
\frac{\partial (\bar{d}(\bar{W}))}{\partial v}
= \sum_{p=1}^{N} (-1)^{|d|d_{1,p-1}}
&\Big(\sum_{i < p : v_{i} = v} (-1)^{d_{1,i-1}(|d|+d_{i,N})} (\prod_{j=i+1}^{p-1} v_{j}) d(v_{p}) (\prod_{j=p+1}^{N} v_{j}) (\prod_{j=1}^{i-1} v_{j})
\\
&+ \sum_{i > p : v_{i} = v} (-1)^{d_{i,N}(|d|+d_{1,i-1})} (\prod_{j=i+1}^{N} v_{j}) (\prod_{j=1}^{p-1} v_{j}) d(v_{p}) (\prod_{j=p+1}^{i-1} v_{j})
\\
&+ \sum_{m \in M_{v_{p}}} \sum_{l \in L_{v_{p}}} (-1)^{(d_{1,p-1}+|r_{l,v_{p}}^{m}|)(|v|+|s_{l,v_{p}}^{m}|+d_{p+1,N})}
s_{l,v_{p}}^{m} (\prod_{j=p+1}^{N} v_{j}) (\prod_{j=1}^{p-1} v_{j}) r_{l,v_{p}}^{m}\Big).
\end{align*}
It is obvious to see that the first two terms of the second member of the previous identity coincide with $(-1)^{|v||d|}$ times the two terms of the last member of
\eqref{eq:dpW}.
It is also clear that the third term of the equation coincides with the second term of the second member of \eqref{eq:difi}, which proves the claim.
The proposition now follows directly from the identity \eqref{eq:difi}, since $\bar{d}(\bar{W}) = 0$.
\qed
\begin{remark}
\label{rem:simpot}
From the proof of the proposition we further see that the ideal $I \subset TV$ generated by the cyclic derivatives of a homogeneous superpotential
$\bar{W} \in HH_{0}(TV)$ is preserved by a homogeneous derivation $d$ if and only if the cyclic derivatives of $\bar{d}(\bar{W})$ belong to $I$.
\end{remark}
For the rest of this subsection we assume that the Yang-Mills algebra $\ym(n,s)^{\Gamma}$ is equivariant and that $V_{1}^{*}$ is an irreducible spin representation of $\so(V_{0},g)$
(\textit{i.e.}, a so-called \emph{minimal supersymmetry} in physical theories).
We have the following direct consequence of the proposition.
Let us consider the collection $d_{c}^{1}$ ($c = 1, \dots, s$) of homogeneous derivations of $TV$ of degree $1$ induced by
\begin{align*}
d_{c}^{1}(x_{i}) &= \sum_{d=1}^{s} \Gamma^{i}_{c,d} z_{d},
\\
d_{c}^{1}(z_{b}) &= \frac{1}{2} \sum_{d=1}^{s} \sum_{i,j=1}^{n} \tilde{\Gamma}^{i,b,d} \Gamma^{j}_{d,c} [x_{i},x_{j}].
\end{align*}
As expected, they are analogous to the supersymmetry transformations that have been considered in supersymmetric gauge theories
(\textit{cf.} \cite{DF99s}, (6.8)).
The following computation is parallel to one already known to physicists long time ago, but we provide it just for completeness (\textit{cf.} \cite{DF99s}, Thm. 6.4).
Applying the derivation $d_{a}^{1}$ to the superpotential $\bar{W}$ given in \eqref{eq:sup}, we see that, on the one hand
\[ d_{c}^{1}\Big(-\frac{1}{4} \sum_{i,j = 1}^{n} [x_{i},x_{j}][x_{i},x_{j}]\Big) = - \sum_{b=1}^{s} \sum_{i,j = 1}^{n} \Gamma^{j}_{c,b} [x_{i},x_{j}][x_{i},z_{b}]. \]
On the other hand, we have
\begin{align*}
d_{c}^{1}\Big(\frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b = 1}^{s} \Gamma_{a,b}^{i} z_{a}[x_{i},z_{b}]\Big)
&= \frac{1}{4} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s}
\Gamma_{a,b}^{i} \tilde{\Gamma}^{j,a,d} \Gamma^{l}_{d,c} [x_{j},x_{l}] [x_{i},z_{b}]
- \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \Gamma^{i}_{c,d} z_{a} [z_{d},z_{b}]
\\
&- \frac{1}{4} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c}
z_{a}[x_{i}, [x_{j},x_{l}]]
\\
&= - \frac{1}{2} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s}
\Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} [x_{i},[x_{j},x_{l}]] z_{a}
- \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \Gamma^{i}_{c,d} z_{a} [z_{d},z_{b}]
\\
&+ \frac{1}{4} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c}
[x_{i}, [x_{j},x_{l}]z_{a}],
\end{align*}
where we have used that $[x_{j},x_{l}][x_{i},z_{a}] = [x_{i},[x_{j},x_{l}]z_{a}] - [x_{i},[x_{j},x_{l}]]z_{a}$ in the first term of the third member.
We now consider the following collection of homogeneous element of $TV$ (for $a,c = 1, \dots, s$)
\begin{align*}
X_{a,c} &= \frac{1}{2} \sum_{i,j,l = 1}^{n} \sum_{b,d = 1}^{s}
\Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} [x_{i},[x_{j},x_{l}]]
= \frac{1}{2} \sum_{i,j,l = 1}^{n} \sum_{b,d = 1}^{s}
\Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} ([x_{l},[x_{j},x_{i}]] + [x_{j},[x_{i},x_{l}]])
\\
&= \frac{1}{2} \sum_{i,j,l = 1}^{n} \sum_{b,d = 1}^{s}
(\Gamma_{c,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,a} + \Gamma_{a,b}^{j} \tilde{\Gamma}^{i,b,d} \Gamma^{l}_{d,c}) [x_{i},[x_{j},x_{l}]]
\\
&= X_{c,a} + \frac{1}{2} \sum_{i,j,l = 1}^{n} \sum_{d = 1}^{s}
(- \sum_{b = 1}^{s} \Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} + 2 \Gamma_{d,c}^{l} \delta_{i,j} \delta_{a,d}) [x_{i},[x_{j},x_{l}]]
\\
&= X_{c,a} - X_{a,c} + S_{a,c},
\end{align*}
where we have used \eqref{eq:relgammasind} in the penultimate equality and we define
\[ S_{a,c} = \sum_{i,l=1}^{n} \Gamma_{a,c}^{l} [x_{i},[x_{i},x_{l}]]. \]
This implies that $2 X_{a,c} = X_{c,a} + S_{a,c}$.
Using that $S_{a,c}$ is symmetric for the interchange of indices, a trivial computation implies that $X_{a,c}$ is also so,
which in turn implies that $X_{a,c} = S_{a,c}$.
As a consequence, we have that
\begin{align*}
d_{c}^{1}\Big(\frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b = 1}^{s} \Gamma_{a,b}^{i} z_{a}[x_{i},z_{b}]\Big)
&= - \sum_{i,l=1}^{n} \sum_{a=1}^{s} \Gamma_{a,c}^{l} [x_{i},[x_{i},x_{l}]] z_{a}
- \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \Gamma^{i}_{c,d} z_{a} [z_{d},z_{b}]
\\
&+ \frac{1}{4} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} [x_{i}, [x_{j},x_{l}]z_{a}]
\\
&= \sum_{i,l=1}^{n} \sum_{a=1}^{s} \Gamma_{a,c}^{l} [x_{i},x_{l}] [x_{i},z_{a}]
- \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \Gamma^{i}_{c,d} z_{a} [z_{d},z_{b}]
\\
&+ \frac{1}{4} \sum_{i,j,l = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \tilde{\Gamma}^{j,b,d} \Gamma^{l}_{d,c} [x_{i}, [x_{j},x_{l}]z_{a}]
- \sum_{i,l=1}^{n} \sum_{a=1}^{s} \Gamma_{a,c}^{l} [x_{i},[x_{i},x_{l}] z_{a}].
\end{align*}
The last two terms belong to $HH_{0}(TV)$, so we may discard them when considering $\bar{d}_{c}^{1}(\bar{W})$.
In consequence, summing up, we obtain that $\bar{d}_{c}^{1}(\bar{W})$ is the class of the element
\[ - \frac{1}{2} \sum_{i = 1}^{n} \sum_{a,b,d = 1}^{s} \Gamma_{a,b}^{i} \Gamma^{i}_{c,d} z_{a} [z_{d},z_{b}], \]
whose cyclic derivatives belong to the ideal generated by the cyclic derivatives of $W$ if and only if it vanishes.
Therefore, using the Jacobi identity and Proposition \ref{prop:simpot} (or, more precisely, Remark \ref{rem:simpot}),
the set of homogeneous derivations $\{ d_{c}^{1} \}_{c=1,\dots,s}$ preserve the ideal $\cl{R(n,s)^{\Gamma}}$ if and only if
the quartic form $\sum_{i=1}^{n} (\Gamma_{a,b}^{i} \Gamma_{c,d}^{i} + \Gamma_{a,c}^{i} \Gamma_{b,d}^{i} + \Gamma_{a,d}^{i} \Gamma_{b,c}^{i})$
(\textit{i.e.} the one given by $z^{*} \mapsto g(\Gamma(z^{*},z^{*}),\Gamma(z^{*},z^{*}))$, for $z^{*} \in V_{1}^{*}$) vanishes.
As explained in \cite{DF99s}, \S 6.1, this happens for $n = 3, 4, 6, 10$.
This also gives a simpler proof of \cite{Mov05}, Prop. 20 (besides the other implication which was not stated there).
\subsection{Another description of the super Yang-Mills algebra}
\label{subsec:anodesc}
Define $\h(n,s)$ to be super Lie algebra generated by the super vector space $U(n,s) = U(n,s)_{0} \oplus U(n,s)_{1}$, where
$U(n,s)_{0} = \mathrm{span}_{k} \cl{q_{2}, \dots, q_{n}, p_{2}, \dots, p_{n}}$ and $U(n,s)_{1} = \mathrm{span}_{k} \cl{z_{1}', \dots, z_{s}'}$,
with the relation space given by $R(n,s)' = k.(\sum_{i=2}^{n} [q_{i},p_{i}] + \frac{1}{2} \sum_{a=1}^{s} [z_{a}',z_{a}'])$,
\textit{i.e.} $\h(n,s) = \f(U(n,s))/\cl{R(n,s)'}$.
The (super) Lie algebra $\a = k.d$, with $d$ even, acts by (even) derivations on $\h(n,s)$ as follows:
\begin{align*}
d(q_{i}) &= p_{i},
\\
d(p_{i}) &= - \sum_{j=2}^{n} [q_{j},[q_{j},q_{i}]] + \frac{1}{2} \sum_{a,b = 1}^{s} \Gamma_{a,b}^{i} [z_{a}',z_{b}'],
\\
d(z_{a}') &= - \sum_{j=2}^{n} \sum_{b=1}^{s} \Gamma_{a,b}^{j} [q_{j},z_{b}'],
\end{align*}
for all $i = 2, \dots, n$ and $a = 1, \dots, s$.
Note that, for the previous action to be well-defined, $\Gamma$ need not be nondegenerate.
We now easily obtain a morphism of super Lie algebras
\[ \psi : \ym(n,s)^{\Gamma} \rightarrow \a \ltimes \h(n,s) \]
given by
\begin{align*}
x_{1} &\mapsto d,
\\
x_{i} &\mapsto q_{i},
\\
z_{a} &\mapsto z_{a}',
\end{align*}
for all $i = 2, \dots, n$ and $a = 1, \dots, s$.
The morphism is well-defined because of the assumption that $\B_{1}^{*}$ is orthonormal with respect to $x_{1}^{*}|_{V_{0}} \circ \Gamma$.
Moreover, it is bijective with inverse given by the morphism of super Lie algebras
\[ \psi^{-1} : \a \ltimes \h(n,s) \rightarrow \ym(n,s)^{\Gamma}, \]
defined as
\begin{align*}
d &\mapsto x_{1},
\\
q_{i} &\mapsto x_{i},
\\
p_{i} &\mapsto [x_{1},x_{i}],
\\
z_{a}' &\mapsto z_{a},
\end{align*}
for all $i = 2, \dots, n$ and $a = 1, \dots, s$.
Therefore, we have that, under the assumption that $\Gamma$ is nondegenerate, $\ym(n,s)^{\Gamma} \simeq \a \ltimes \h(n,s)$,
which further yields that $\YM(n,s)^{\Gamma} \simeq \U(\a) \# H(n,s)$, where $H(n,s) = \U(\h(n,s))$ (\textit{cf.}~\cite{MS06}, Prop. 11).
In particular, we may regard $\h(n,s)$ canonically included in $\ym(n,s)^{\Gamma}$ (using the morphism $\psi^{-1}$).
We would also like to point out that all previous isomorphisms also hold if we consider $\ym(n,s)^{\Gamma}$ as a graded Lie algebra with generators $x_{i}$ in degree $2$ and generators $z_{a}$ in degree $3$,
$\a$ as the free graded Lie algebra with generator $d$ of degree $2$, and $\h(n,s)$ as a graded Lie algebra with generators $q_{i}$ of degree $2$,
$p_{i}$ of degree $4$ and $z_{a}'$ of degree $3$, with the given relation (of degree $6$).
\section{Several homological computations}
\label{sec:hom}
\subsection{A Koszul-like projective resolution}
\label{subsec:projres}
In this subsection we shall provide the minimal projective resolution of the left $\YM(n,s)^{\Gamma}$-module $k$,
when $\YM(n,s)^{\Gamma}$ is considered as a graded algebra, where we suppose that $(n,s) \in (\NN \times \NN_{0}) \setminus \{ (1,0), (1,1) \}$.
We have excluded the case $(n,s) = (1,0)$, because in this case the super Yang-Mills algebra is a polynomial algebra in one (even) variable.
Using ideas analogous to the case of Koszul algebras, we shall construct from the previous resolution the minimal projective resolution of the $\YM(n,s)^{\Gamma}$-bimodule $\YM(n,s)^{\Gamma}$,
and derive several consequences from the explicit shape of the previous resolutions.
Although some of the results of the first part of this subsection are mentioned in \cite{MS06}, our aim here is to give detailed proofs of the results that we will need later.
Regarding $\YM(n,s)^{\Gamma}$ as a nonnegatively graded connected algebra with the grading stated in Subsection \ref{subsec:defi},
and since $k$ is a bounded below graded module over $\YM(n,s)^{\Gamma}$,
we know that there exists a minimal projective resolution of $k$ over $\YM(n,s)^{\Gamma}$, which is of the form
$K_{\bullet}' = \YM(n,s)^{\Gamma} \otimes_{k} \Tor_{\bullet}^{\YM(n,s)^{\Gamma}}(k,k)$
(see \cite{Ber08}, Th\'eo. 1.11, Prop. 2.3).
In order to obtain this minimal projective resolution, we shall proceed as follows.
First, we will compute the homology groups $H_{\bullet}(\h(n,s),k)$ together with their action of $\a$.
From the Hochschild-Serre spectral sequence $E_{p,q}^{2} = H_{p}(\a,H_{q}(\h(n,s),k)) \Rightarrow H_{p+q}(\ym(n,s)^{\Gamma},k)$, we shall obtain
the homology groups $H_{\bullet}(\ym(n,s)^{\Gamma},k)$.
These computations will in turn give the minimal projective resolution of $k$ over $\YM(n,s)^{\Gamma}$.
We remark that all algebras $\a$, $\h(n,s)$ and $\ym(n,s)^{\Gamma}$ in these computations are considered as graded Lie algebras with the grading explained at the end of Subsections \ref{subsec:anodesc}.
We first compute the homology groups $H_{\bullet}(\h(n,s),k)$ together with their action of $\a$.
The minimal projective resolution of $k$ as a module over $H(n,s) = \U(\h(n,s))$ is of the form
\begin{equation}
\label{eq:hns}
0 \rightarrow H(n,s) \otimes R(n,s)' \overset{d_{2}'}{\rightarrow} H(n,s) \otimes U(n,s) \overset{d_{1}'}{\rightarrow}
H(n,s) \overset{\epsilon_{\h(n,s)}}{\rightarrow} k \rightarrow 0,
\end{equation}
where we recall that $d_{2}'$ is given by the restriction of the map $H(n,s) \otimes U(n,s)^{\otimes 2} \rightarrow H(n,s) \otimes U(n,s)$
of the form $z \otimes (v \otimes w) \mapsto z.v \otimes w$, for $z \in H(n,s)$ and $v, w \in U(n,s)$, and
$d_{1}'$ is the restriction of the multiplication of $H(n,s)$.
Indeed, if $H(n,s)$ is considered to be generated in degree one, \textit{i.e.} $U(n,s)$ is seen to be concentrated in degree $1$, $H(n,s)$ becomes a quadratic algebra with only one relation, therefore Koszul
(see \cite{Fr75}, Thm. 3).
moreover, the minimal projective resolution of the $H(n,s)$-module $k$ is given by \eqref{eq:hns}, because the map $d_{2}'$ is injective.
To prove this last statement we may note that, if $n \geq 2$ then $\U(\ym(n,s)^{\Gamma}) \supset \U(\ym(n))$ is a free extension of algebras with $\U(\ym(n))$ a
domain, and if $n = 1$ and $s \geq 2$, it follows from a simple computation using an explicit basis of the algebra $\U(\h(1,s))$ given by the Diamond Lemma (see \cite{Berg78}).
Since the resolution \eqref{eq:hns} also holds for the special grading of $H(n,s)$, \textit{i.e.} when $\deg(q_{i}) = 2$, $\deg(p_{i}) = 4$, and $\deg(z_{a}') = 3$, for all $i = 2, \dots, n$ and $a = 1, \dots, s$,
the claim follows.
In particular we have canonical isomorphisms $H_{0}(\h(n,s),k) \simeq k$, $H_{1}(\h(n,s),k) \simeq U(n,s)$, $H_{2}(\h(n,s),k) \simeq R(n,s)'$
and the other homology groups vanish.
In order to obtain the action of $\a$ on the homology groups we compare the previous resolution with the Chevalley-Eilenberg resolution of $k$ over
the graded Lie algebra $\ym(n,s)^{\Gamma}$, seen as a resolution of $k$ as $\h(n,s)$-modules.
We recall that, for a graded (or super) Lie algebra $\g$, the Chevalley-Eilenberg resolution of the trivial $\g$-module $k$ is of the form $(\U(\g) \otimes \Lambda^{\bullet} \g, d^{\mathrm{CE}}_{\bullet})$,
where we recall that the exterior tensor products are taken in the graded (or super) sense, and the differential is given by (\textit{cf.} \cite{Ta95}, Section 1)
\begin{multline*}
d_{n}^{\mathrm{CE}}(z \otimes y_{1} \wedge \dots \wedge y_{n}) = \sum_{i=1}^{n} (-1)^{|y_{i}|(|y_{1}|+\dots+|y_{i-1}|) + (i-1)} z y_{i} \otimes y_{1} \wedge \dots \wedge \hat{y}_{i} \wedge \dots \wedge y_{n}
\\
+ \sum_{1 \leq i < j \leq n} (-1)^{|y_{i}|(|y_{1}|+\dots+|y_{i-1}|) + |y_{j}|(|y_{1}|+\dots+|y_{j-1}|) + |y_{i}||y_{j}| + (i+j)}
z \otimes [y_{i},y_{j}] \wedge y_{1} \wedge \dots \wedge \hat{y}_{i} \wedge \dots \wedge \hat{y}_{j} \wedge \dots \wedge y_{n},
\end{multline*}
for $y_{1}, \dots y_{n} \in \g$ homogeneous elements and $z \in \U(\g)$.
We have the following comparison morphism of resolutions of projective $H(n,s)$-modules of $k$:
\[
\xymatrix
{
\dots
\ar[r]
&
0
\ar[r]
\ar[d]
&
H \otimes R'
\ar[r]^{d_{2}'}
\ar[d]^{\mathrm{can}}
&
H \otimes U
\ar[r]^{d_{1}'}
\ar[d]^{\mathrm{inc}}
&
H
\ar[r]^{\epsilon_{\h}}
\ar[d]^{\mathrm{inc}}
&
k
\ar[r]
\ar@{=}[d]
&
0
\\
\dots
\ar[r]^-{d_{4}^{CE}}
&
\YM \otimes \Lambda^{3} \ym
\ar[r]^-{d_{3}^{CE}}
&
\YM \otimes \Lambda^{2} \ym
\ar[r]^-{d_{2}^{CE}}
&
\YM \otimes \ym
\ar[r]^-{d_{1}^{CE}}
&
\YM
\ar[r]^{\epsilon_{\ym}}
&
k
\ar[r]
&
0
}
\]
where we have omitted the indices $(n,s)$ and $\Gamma$ for simplicity,
\[ \mathrm{can}(z \otimes (\sum_{i=2}^{n} [q_{i},p_{i}] + \frac{1}{2} \sum_{a=1}^{s} [z_{a}',z_{a}']))
= z \otimes (\sum_{i=2}^{n} x_{i} \wedge [x_{1},x_{i}] + \frac{1}{2} \sum_{a=1}^{s} z_{a} \wedge z_{a}), \]
and we remark that the exterior power and the wedge product here are in the supersymmetric sense.
Applying the functor $(\place)_{\h(n,s)}$ to both resolutions, we see that, under the previous comparison morphisms,
\begin{itemize}
\item[(i)] $\bar{1} \in (\YM(n,s)^{\Gamma})_{\h(n,s)}$ corresponds to $1 \in k \simeq H_{0}(\h(n,s))$,
\item[(ii)] $\overline{1 \otimes x_{2}}, \dots, \overline{1 \otimes x_{n}}, \overline{1 \otimes [x_{1},x_{2}]}, \dots, \overline{1 \otimes [x_{1},x_{n}]},
\overline{1 \otimes z_{1}}, \dots, \overline{1 \otimes z_{s}} \in (\YM(n,s)^{\Gamma} \otimes \ym(n,s))_{\h(n,s)}$
correspond to $q_{2}, \dots, q_{n}, p_{2}, \dots, p_{n}, z_{1}', \dots, z_{s}' \in U(n,s) \simeq H_{1}(\h(n,s),k)$,
\item[(ii)] $\overline{1 \otimes (\sum_{i=2}^{n} x_{i} \wedge [x_{1} , x_{i}] + (1/2) \sum_{a=1}^{s} z_{a} \wedge z_{a})}
\in (\YM(n,s)^{\Gamma} \otimes \Lambda^{2} \ym(n,s))_{\h(n,s)}$
corresponds to the relation element given by $\sum_{i=2}^{n} [q_{i},p_{i}] + (1/2) \sum_{a=1}^{s} [z_{a}',z_{a}'] \in R(n,s)' \simeq H_{2}(\h(n,s))$.
\end{itemize}
The action of $\a = k.d$ on the homology groups is induced by the action of $x_{1}$ on the tensor component $\YM(n,s)^{\Gamma}$
of the modules $\YM(n,s)^{\Gamma} \otimes \Lambda^{\bullet} \ym(n,s)^{\Gamma}$ of the Chevalley-Eilenberg resolution.
This immediately implies that the action of $x_{1}$ on $H_{0}(\h(n,s),k) \simeq k$ is trivial, for $x_{1} = d^{CE}_{1} (1 \otimes x_{1})$ is a boundary.
Concerning the action of $\a$ on $H_{1}(\h(n,s),k)$, we have the identity $x_{1} . q_{i} = p_{i}$, for all $i = 2, \dots, n$, since
\[ x_{1} \otimes x_{i} = 1 \otimes [x_{1},x_{i}] + x_{i} \otimes x_{1} + d^{CE}_{2}(1 \otimes (x_{1} \wedge x_{i})), \]
and $\overline{x_{i} \otimes x_{1}} \in (\YM(n,s)^{\Gamma} \otimes \ym(n,s)^{\Gamma})_{\h(n,s)}$ vanishes.
Moreover, the action of $x_{1}$ on both $p_{i}$ and $z_{a}$ is zero, for $i = 2, \dots, n$ and for $a = 1, \dots, s$, respectively.
To prove the first half the statement, note that
\begin{align*}
x_{1} \otimes [x_{1},x_{i}] &= [x_{1},x_{i}] \otimes x_{1} + 1 \otimes [x_{1},[x_{1},x_{i}]] + d^{CE}_{2}(1 \otimes (x_{1} \wedge [x_{1},x_{i}]))
\\
&= [x_{1},x_{i}] \otimes x_{1} - \sum_{j=2}^{n} 1 \otimes [x_{j},[x_{j},x_{i}]]
+ \frac{1}{2} \sum_{a,b=1}^{s} \Gamma_{a,b}^{i} 1 \otimes [z_{a},z_{b}]
+ d^{CE}_{2}(1 \otimes (x_{1} \wedge [x_{1},x_{i}]))
\\
&= [x_{1},x_{i}] \otimes x_{1} + d^{CE}_{2}(\sum_{j=2}^{n} 1 \otimes x_{j} \wedge [x_{j},x_{i}])
- \sum_{j=2}^{n} (x_{j} \otimes [x_{j},x_{i}] - [x_{j},x_{i}] \otimes x_{j})
\\
& - d^{CE}_{2}\Big(\frac{1}{2} \sum_{a,b=1}^{s} \Gamma_{a,b}^{i} 1 \otimes z_{a} \wedge z_{b}\Big)
+ \sum_{a,b=1}^{s} \Gamma_{a,b}^{i} z_{a} \otimes z_{b}
+ d^{CE}_{2}(1 \otimes (x_{1} \wedge [x_{1},x_{i}])),
\end{align*}
and the elements $\overline{[x_{1},x_{i}] \otimes x_{1}}$, $\overline{x_{j} \otimes [x_{j},x_{i}] - [x_{j},x_{i}] \otimes x_{j}}$
and $\overline{z_{a} \otimes z_{b}}$ of the space $(\YM(n,s)^{\Gamma} \otimes \ym(n,s)^{\Gamma})_{\h(n,s)}$ vanish.
The second half of the statement follows from
\[ x_{1} \otimes z_{a} = 1 \otimes [x_{1},z_{a}] + z_{a} \otimes x_{1} + d^{CE}_{2}(1 \otimes (x_{1} \wedge z_{a})) \]
and the fact that $\overline{z_{a} \otimes x_{1}} \in (\YM(n,s)^{\Gamma} \otimes \ym(n,s)^{\Gamma})_{\h(n,s)}$ vanishes.
Finally, the action of $x_{1}$ on $H_{2}(\h(n,s),k) \simeq R(n,s)'$ is trivial, which can be proved as follows.
First, note that
\begin{align*}
x_{1} \otimes \Big(\sum_{i=2}^{n} x_{i} \wedge [x_{1} , x_{i}] + \frac{1}{2} \sum_{a=1}^{s} z_{a} \wedge z_{a}\Big)
&= d^{CE}_{3}(1 \otimes (x_{1} \wedge x_{i} \wedge [x_{1},x_{i}]))
+ \sum_{i=2}^{n} (x_{i} \otimes x_{1} \wedge [x_{1},x_{i}] - [x_{1},x_{i}] \otimes x_{1} \wedge x_{i})
\\
& - \sum_{i=2}^{n} 1 \otimes x_{1} \wedge [x_{i},[x_{1},x_{i}]]
+ \sum_{i=2}^{n} 1 \otimes x_{i} \wedge [x_{1},[x_{1},x_{i}]]
+ \sum_{a=1}^{s} z_{a} \otimes x_{1} \wedge z_{a}
\\
& + d^{CE}_{3}\Big(\frac{1}{2} \sum_{a=1}^{s} 1 \otimes x_{1} \wedge z_{a} \wedge z_{a}\Big)
+ \sum_{a=1}^{s} 1 \otimes [x_{1},z_{a}] \wedge z_{a}
- \frac{1}{2} \sum_{a=1}^{s} 1 \otimes x_{1} \wedge [z_{a}, z_{a}]
\\
&= d^{CE}_{3}(1 \otimes (x_{1} \wedge x_{i} \wedge [x_{1},x_{i}]))
+ \sum_{i=2}^{n} (x_{i} \otimes x_{1} \wedge [x_{1},x_{i}] - [x_{1},x_{i}] \otimes x_{1} \wedge x_{i})
\\
& + \sum_{i=2}^{n} 1 \otimes x_{i} \wedge [x_{1},[x_{1},x_{i}]]
+ \sum_{a=1}^{s} z_{a} \otimes x_{1} \wedge z_{a}
+ d^{CE}_{3}\Big(\frac{1}{2} \sum_{a=1}^{s} 1 \otimes x_{1} \wedge z_{a} \wedge z_{a}\Big)
\\
& + \sum_{a=1}^{s} 1 \otimes [x_{1},z_{a}] \wedge z_{a},
\end{align*}
where we have simplified the last member using the first relation $r_{0,1}$ of \eqref{eq:relort}.
Since the elements $\overline{x_{i} \otimes x_{1} \wedge [x_{1},x_{i}] - [x_{1},x_{i}] \otimes x_{1} \wedge x_{i}}$ and
$\overline{z_{a} \otimes x_{1} \wedge z_{a}}$ of $(\YM(n,s)^{\Gamma} \otimes \Lambda^{2} \ym(n,s)^{\Gamma})_{\h(n,s)}$ vanish,
it suffices to prove that the element of $(\YM(n,s)^{\Gamma} \otimes \Lambda^{2} \ym(n,s)^{\Gamma})_{\h(n,s)}$ induced by
\[ \sum_{i=2}^{n} 1 \otimes x_{i} \wedge [x_{1},[x_{1},x_{i}]] + \sum_{a=1}^{s} 1 \otimes [x_{1},z_{a}] \wedge z_{a} \]
also vanish, which follows from
\begin{align*}
\sum_{i=2}^{n} 1 \otimes x_{i} \wedge [x_{1},[x_{1},x_{i}]] + \sum_{a=1}^{s} 1 \otimes [x_{1},z_{a}] \wedge z_{a}
&= - \sum_{i,j=2}^{n} 1 \otimes x_{i} \wedge [x_{j},[x_{j},x_{i}]]
+ \frac{1}{2} \sum_{i=2}^{n} \sum_{a,b=1}^{s} \Gamma^{i}_{a,b} 1 \otimes x_{i} \wedge [z_{a},z_{b}]
\\
& - \sum_{a,b=1}^{s} \sum_{i=2}^{s} \Gamma^{i}_{a,b} 1 \otimes [x_{i},z_{b}] \wedge z_{a}
\\
&= \frac{1}{2} d^{CE}_{3}\Big(\sum_{i,j=2}^{n} 1 \otimes x_{i} \wedge x_{j} \wedge [x_{i},x_{j}]
+ \sum_{i=2}^{n} \sum_{a,b=1}^{s} \Gamma^{i}_{a,b} 1 \otimes x_{i} \wedge z_{b} \wedge z_{a}\Big)
\\
& - \frac{1}{2} \sum_{i=2}^{n} \sum_{a,b=1}^{s} \Gamma^{i}_{a,b} (x_{i} \otimes z_{b} \wedge z_{a} - 2 z_{b} \otimes x_{i} \wedge z_{a})
- \sum_{i,j=2}^{n} x_{i} \otimes x_{j} \wedge [x_{i},x_{j}]
\\
&- \frac{1}{2} \sum_{i,j=2}^{n} [x_{i},x_{j}] \otimes x_{i} \wedge x_{j}.
\end{align*}
We now proceed to compute the homology groups $H_{\bullet}(\a,H_{\bullet}(\h(n,s),k))$.
In this case, the Chevalley-Eilenberg resolution is of the form
\[ 0 \rightarrow \U(\a) \otimes k.d \rightarrow \U(\a) \overset{\epsilon_{\a}}{\rightarrow} k \rightarrow 0, \]
where the first nonzero morphism is given by the restriction of the multiplication map of $\U(\a)$.
In particular, $H_{0}(\a,k) \simeq k$, $H_{1}(\a,k) \simeq k[-2]$, $H_{0}(\a,R(n,s)') \simeq k[-6]$, $H_{1}(\a,R(n,s)') \simeq k[-8]$, and zero otherwise.
Moreover, the previous resolution tells us that
\begin{align*}
H_{0}(\a,U(n,s)) &\simeq \mathrm{span}_{k} \cl{q_{2}, \dots, q_{n}, z_{1}', \dots, z_{s}'},
\\
H_{1}(\a,U(n,s)) &\simeq \mathrm{span}_{k} \cl{p_{2}, \dots, p_{n}, z_{1}', \dots, z_{s}'}[-2],
\end{align*}
and the other homology groups vanish.
Now, since the Hochschild-Serre spectral sequence $E_{p,q}^{2} = H_{p}(\a,H_{q}(\h(n,s),k)) \Rightarrow H_{p+q}(\ym(n,s)^{\Gamma},k)$
is concentrated in columns $p = 0 ,1$, it follows that $H_{n}(\ym(n,s)^{\Gamma},k) \simeq H_{0}(\a,H_{n}(\h(n,s),k)) \oplus H_{1}(\a,H_{n-1}(\h(n,s),k))$,
for all $n \in \NN_{0}$ (see \cite{Rot09}, Cor. 10.29).
Hence,
\begin{equation}
\label{eq:homo}
H_{\bullet}(\ym(n,s)^{\Gamma},k) \simeq \begin{cases}
k, &\text{if $\bullet = 0$,}
\\
V(n,s), &\text{if $\bullet = 1$,}
\\
R(n,s)^{\Gamma}, &\text{if $\bullet = 2$,}
\\
k.\omega, &\text{if $\bullet = 3$,}
\\
0, &\text{else,}
\end{cases}
\end{equation}
where $\omega = \sum_{i=1}^{n} x_{i} \otimes r_{0,i} + \sum_{a=1}^{s} z_{a} \otimes r_{1,a} \in (TV)_{8}$.
Note that $k.\omega \simeq k[-8]$ and $R(n,s)^{\Gamma} \simeq V(n,s)^{*}[-8]$ in the category of graded $\so(n)$-modules provided with morphisms of degree zero.
Also note that $\omega \in (V \otimes R) \cap (R \otimes V)$, because of the identity
\[ \sum_{i=1}^{n} [x_{i}, r_{0,i}] + \sum_{a=1}^{s} [z_{a}, r_{1,a}] = 0 \]
in the tensor algebra $TV$.
Moreover, we have the following:
\begin{proposition}
\label{prop:projres}
Let $(n,s) \in (\NN \times \NN_{0}) \setminus \{ (1,0), (1,1) \}$.
The minimal projective resolution $(K_{\bullet}'(\ym(n,s)^{\Gamma}),b_{\bullet}')$ of the $\YM(n,s)^{\Gamma}$-module $k$ is given by
\begin{equation}
\label{eq:comp}
0 \rightarrow \YM(n,s)^{\Gamma}[-8] \overset{b_{3}'}{\rightarrow} \YM(n,s)^{\Gamma} \otimes R(n,s)^{\Gamma} \overset{b_{2}'}{\rightarrow}
\YM(n,s)^{\Gamma} \otimes V(n,s) \overset{b_{1}'}{\rightarrow} \YM(n,s)^{\Gamma} \overset{b_{0}'}{\rightarrow} k \rightarrow 0,
\end{equation}
with differential
\begin{equation}
\label{eq:dif}
\begin{split}
b_{3}'(z) &= \sum_{i=1}^{n} z x_{i} \otimes r_{0,i} + \sum_{a=1}^{s} z z_{a} \otimes r_{1,a},
\\
b_{2}'(z \otimes r_{0,i}) &= \sum_{j=1}^{n} (z x_{j}^{2} \otimes x_{i} - 2 z x_{j} x_{i} \otimes x_{j} + z x_{i} x_{j} \otimes x_{j})
- \sum_{a,b=1}^{s} \Gamma_{a,b}^{i} z z_{a} \otimes z_{b},
\\
b_{2}'(z \otimes r_{1,a}) &= \sum_{i=1}^{n} \sum_{b=1}^{s} \Gamma^{i}_{a,b} (z x_{i} \otimes z_{b} - z z_{b} \otimes x_{i}),
\\
b_{1}'(z \otimes x_{i}) &= z x_{i},
\\
b_{1}'(z \otimes z_{a}) &= z z_{a},
\\
b_{0}'(z) &= \epsilon_{\ym(n,s)^{\Gamma}}(z).
\end{split}
\end{equation}
\end{proposition}
\noindent\textbf{Proof.}
As stated above, the $\YM(n,s)^{\Gamma}$-modules $K_{\bullet}'(\ym(n,s)^{\Gamma},k)$ giving the minimal projective resolution of $k$ are
of the form $\YM(n,s)^{\Gamma} \otimes_{k} \Tor^{\YM(n,s)^{\Gamma}}_{\bullet}(k,k)$.
Since $H_{\bullet}(\ym(n,s)^{\Gamma},k) \simeq \Tor^{\YM(n,s)^{\Gamma}}_{\bullet}(k,k)$, the previous homological computations tell us that the modules involved
in \eqref{eq:comp} are correct and we only need to prove that the claimed differential provides a resolution of $k$.
It is obvious to see that this gives a complex, \textit{i.e.} $b_{i}' \circ b_{i+1}' = 0$, for $i = 0, 1, 2$.
A similar argument to the one given after Lemma 2.3 in \cite{BG06} shows that
\[ \YM(n,s)^{\Gamma} \otimes R(n,s)^{\Gamma} \overset{b_{2}'}{\rightarrow}
\YM(n,s)^{\Gamma} \otimes V(n,s) \overset{b_{1}'}{\rightarrow} \YM(n,s)^{\Gamma} \overset{b_{0}'}{\rightarrow} k \rightarrow 0, \]
is exact.
Indeed, let $A = TV/\cl{R}$ be a connected graded algebra, where $V = \oplus_{i \in \NN} V_{i}$ and $R = \oplus_{i \in \NN_{\geq 2}} R_{i} \subseteq T^{+}V = \oplus_{i \in \NN} V^{\otimes i}$ are
nonnegatively graded vector spaces and we further suppose $R \cap ((T^{+}V) R (TV) + (TV) R (T^{+}V)) = 0$, in order to avoid ambiguities.
In this case, we consider the complex
\[ A \otimes R \overset{b_{2}'}{\rightarrow} A \otimes V \overset{b_{1}'}{\rightarrow} A \overset{b_{0}'}{\rightarrow} k \rightarrow 0, \]
where $b_{0}'$ is the augmentation of $A$, $b_{1}'$ is the restriction of the multiplication of $A$, and $b_{2}'$ is the $A$-linear extension
of the morphism $R \rightarrow TV \otimes V \rightarrow A \otimes V$, given by the composition of the inclusion of $R$ in $T^{+}V \simeq TV \otimes V$ with the canonical surjection.
Since the zeroth degree component of $b_{1}'$ is zero, whereas the $m$-th degree component ($m \geq 1$) is the surjective map
\[ \sum_{i \in \NN} \frac{(TV)_{m-i} \otimes V_{i}}{\cl{R}_{m-i} V_{i}} \rightarrow \frac{(TV)_{m}}{\cl{R}_{m}}, \]
where $\cl{R}_{m}$ is the component of degree $m$ of the ideal generated by $R$ in $TV$,
it follows that $\Ker(b_{0}') = \mathrm{Im}(b_{1}')$.
Moreover, taking into account that the $m$-th degree component of $b_{2}'$ is the map
\[ \frac{\sum_{i \in \NN} (TV)_{m-i} R_{i}}{\sum_{i \in \NN} \cl{R}_{m-i} R_{i}}
\rightarrow \frac{\sum_{j \in \NN} (TV)_{m-j} V_{j}}{\sum_{j \in \NN} \cl{R}_{m-j} V_{j}}, \]
then
\[ \mathrm{Im}(b_{2}')_{m}
= \frac{\sum_{i \in \NN} (TV)_{m-i} R_{i} + \sum_{j \in \NN} \cl{R}_{m-j} V_{j}}{\sum_{j \in \NN} \cl{R}_{m-j} V_{j}}, \]
which coincides with
\[ \Ker(b_{1}')_{m} = \frac{\cl{R}_{m}}{\sum_{j \in \NN} \cl{R}_{m-j} V_{j}}. \]
It is still left to show that $b_{3}'$ is injective and that $\mathrm{Im}(b_{3}') = \Ker(b_{2}')$.
Let $z \in \YM(n,s)^{\Gamma}$ such that $b_{3}'(z) = 0$, \textit{i.e.} such that $z x_{i} = 0$ and $z z_{a} = 0$ for all $i = 1, \dots, n$ and $a = 1, \dots, s$.
Since $\YM(n,s)^{\Gamma}$ is a free right module over the integral domain $\U(\ym(n,s)^{\Gamma}_{0})$, we conclude that $z = 0$, so $b_{3}'$ is injective.
In oder to prove that $\mathrm{Im}(b_{3}') = \Ker(b_{2}')$, we proceed as follows.
Since the third component of the minimal projective resolution of $k$ is $\YM(n,s)^{\Gamma} \otimes k[-8]$,
the kernel of $b_{2}'$ should be of the form $i : \YM(n,s)^{\Gamma} \otimes k[-8] \rightarrow \YM(n,s)^{\Gamma} \otimes R(n,s)^{\Gamma}$.
Hence, we have the commutative diagram
\[
\xymatrix@C-10pt
{
\YM(n,s)^{\Gamma} \otimes k[-8]
\ar[rd]^{b_{3}'}
\ar@{-->}[dd]^{\bar{i}}
&
&
\\
&
\YM(n,s)^{\Gamma} \otimes R(n,s)^{\Gamma}
\ar[r]^{b_{2}'}
&
\YM(n,s)^{\Gamma} \otimes V(n,s)
\\
\YM(n,s)^{\Gamma} \otimes k[-8]
\ar[ru]^{i}
&
&
}
\]
Since $b_{3}'$ is injective, $\bar{i}$ is also so, which in turn implies that it is an isomorphism, for it is an injective endomorphism (of degree zero)
of a locally finite dimensional graded vector space.
Hence, $\mathrm{Im}(b_{3}') = \Ker(b_{2}')$, as was to be proved.
\qed
\begin{remark}
Note that for $(n,s) = (1,1)$ the previous complex does not yield a resolution of $k$, since $H_{\bullet}(\h(1,1),k) \simeq k$, for $\bullet \in \NN_{0}$.
\end{remark}
There is an analogous free resolution of right $\YM(n,s)^{\Gamma}$-modules of the trivial module $k$, for $(n,s) \neq (1,0), (1,1)$, which is of the form
\begin{equation}
\label{eq:comp*}
0 \rightarrow \YM(n,s)^{\Gamma}[-8] \overset{b_{3}''}{\rightarrow} R(n,s)^{\Gamma} \otimes \YM(n,s)^{\Gamma} \overset{b_{2}''}{\rightarrow}
V(n,s) \otimes \YM(n,s)^{\Gamma} \overset{b_{1}''}{\rightarrow} \YM(n,s)^{\Gamma} \overset{b_{0}'}{\rightarrow} k \rightarrow 0,
\end{equation}
with differential given by $b_{0}'' = \epsilon_{\ym(n,s)^{\Gamma}}$ and
\begin{equation}
\label{eq:dif*}
\begin{split}
b_{3}''(z) &= \sum_{i=1}^{n} r_{0,i} \otimes x_{i} z - \sum_{a=1}^{s} r_{1,a} \otimes z_{a} z,
\\
b_{2}''(r_{0,a} \otimes z) &= \sum_{j=1}^{n} (x_{i} \otimes x_{j}^{2} z - 2 x_{j} \otimes x_{i} x_{j} z + x_{j} \otimes x_{j} x_{i} z)
- \sum_{a,b=1}^{s} \Gamma_{a,b}^{i} z_{a} \otimes z_{b} z,
\\
b_{2}''(r_{1,a} \otimes z) &= \sum_{i=1}^{n} \sum_{b=1}^{s} \Gamma^{i}_{a,b} (x_{i} \otimes z_{b} z - z_{b} \otimes x_{i} z),
\\
b_{1}''(x_{i} \otimes z) &= x_{i} z,
\\
b_{1}''(z_{a} \otimes z) &= z_{a} z.
\end{split}
\end{equation}
We shall denote this resolution by $(K_{\bullet}''(\ym(n,s)^{\Gamma}),b_{\bullet}'')$.
We now recall that a nonnegatively graded connected algebra $A$ is called \emph{(left) AS-regular of dimension $d$ and of Gorenstein parameter $l$}
if it has finite (left) global dimension $d$ and it satisfies that
\[ \mathcal{E}xt^{i}_{A} (k, A) \simeq \begin{cases}
k[l], &\text{if $i = d$},
\\
0, &\text{else}.
\end{cases}
\]
We point out that no noetherianity assumption on $A$ is required (see \cite{MM11}, Def. 1.1).
Since we shall be dealing with graded Hopf algebras, both left and right definitions of AS-regular coincide in this case,
so we will omit the reference to the side.
Using the fact that $\mathcal{H}om_{\YM(n,s)^{\Gamma}}(\YM(n,s)^{\Gamma}[i],\place) \simeq (\place)[-i]$, that
\begin{align*}
\YM(n,s)^{\Gamma} \otimes V(n,s) &\simeq (\YM(n,s)^{\Gamma})^{n}[-2] \oplus (\YM(n,s)^{\Gamma})^{s}[-3],
\\
\YM(n,s)^{\Gamma} \otimes R(n,s)^{\Gamma} &\simeq (\YM(n,s)^{\Gamma})^{n}[-6] \oplus (\YM(n,s)^{\Gamma})^{s}[-5],
\end{align*}
and an elementary computation involving the differentials,
one sees that the functor $\mathcal{H}om_{\YM(n,s)^{\Gamma}}(\place,\YM(n,s)^{\Gamma})$ sends the resolution
$(K'_{\bullet}(\ym(n,s)^{\Gamma}),b_{\bullet}')$
of left $\YM(n,s)^{\Gamma}$-modules of $k$ to the shift $(K''_{\bullet}(\ym(n,s)^{\Gamma}),b_{\bullet}'')[8]$.
We remark that the we are always using Koszul's sign rule, and, in particular,
$X \simeq X^{**}$, via the map $x \mapsto (f \mapsto (-1)^{|x||f|}) f(x)$;
$X \otimes Y \simeq Y \otimes X$, via $x \otimes y \mapsto (-1)^{|y||x|} y \otimes x$, and
$Y \otimes X^{*} \simeq \mathcal{H}om_{k}(X,Y)$, via $y \otimes f \mapsto (x \mapsto y f(x))$,
for (finite dimensional) graded (resp. super) vector spaces $X$ and $Y$, and homogeneous elements $x \in X$, $y \in Y$ and $f \in X^{*}$.
This tells us that $\YM(n,s)^{\Gamma}$ is AS-regular with Gorenstein parameter $8$, in the sense of \cite{MM11}, for $(n,s) \neq (1,1)$.
As a corollary of the description of the left and right $\YM(n,s)^{\Gamma}$-module resolutions of $k$,
we may obtain the minimal projective resolution of the $\YM(n,s)^{\Gamma}$-bimodule $\YM(n,s)^{\Gamma}$, for $(n,s) \neq (1,0), (1,1)$.
Consider first the free graded $\YM(n,s)^{\Gamma}$-bimodule
$K_{\bullet}(\ym(n,s)^{\Gamma}) = \YM(n,s)^{\Gamma} \otimes H_{\bullet}(\ym(n,s)^{\Gamma},k) \otimes \YM(n,s)^{\Gamma}$, for $\bullet \geq 0$,
given with the obvious action, where $H_{\bullet}(\ym(n,s)^{\Gamma},k)$ is identified to a subspace of $TV(n,s)$ following \eqref{eq:homo}.
Then, we define on it the differential given by $b_{1} = b_{1}' \otimes 1_{\YM(n,s)^{\Gamma}} - 1_{\YM(n,s)^{\Gamma}} \otimes b_{1}''$,
$b_{3} = b_{3}' \otimes 1_{\YM(n,s)^{\Gamma}} - 1_{\YM(n,s)^{\Gamma}} \otimes b_{3}''$, and
\[ b_{2} (y \otimes y_{1} \dots y_{r} \otimes y') = \sum_{i=1}^{r} y y_{1} \dots y_{i-1} \otimes y_{i} \otimes y_{i+1} \dots y_{r} y', \]
where $y, y' \in \YM(n,s)^{\Gamma}$, $y_{1} \dots y_{r} \in R(n,s)^{\Gamma}$, for $y_{i} \in V(n,s)$, $i = 1, \dots, r$.
Set $b_{\bullet} = 0$, for $\bullet \geq 4$.
It is easily proved to be a complex.
If we define $b_{0} : K_{\bullet}(\ym(n,s)^{\Gamma}) \rightarrow \YM(n,s)^{\Gamma}$ given by the multiplication, we obtain an augmented complex.
Moreover, taking into account the isomorphism
$(K_{\bullet}(\ym(n,s)^{\Gamma}), b_{\bullet}) \otimes_{\YM(n,s)^{\Gamma}} k \simeq (K'_{\bullet}(\ym(n,s)^{\Gamma}),b_{\bullet}')$
of complexes of left $\YM(n,s)^{\Gamma}$-modules,
\cite{BM06}, Prop. 4.1, tells us that the bimodule complex $(K_{\bullet}(\ym(n,s)^{\Gamma}), b_{\bullet})$ is in fact a resolution of $\YM(n,s)^{\Gamma}$.
Let us recall that a locally finite dimensional nonnegatively graded and connected algebra $A$ is called
\emph{(left) graded Calabi-Yau of dimension $d$ and of parameter $l$}
if it has a finite resolution composed of finitely generated projective bimodules, finite global dimension $d$, and
satisfies that (\textit{cf.} \cite{BT07}, Def. 4.2)
\[ \mathcal{E}xt^{i}_{A^{e}} (A, A^{e}) \simeq \begin{cases}
A[l], &\text{if $i = d$},
\\
0, &\text{else},
\end{cases}
\]
in the category of (right) $A^{e}$-modules.
Again, taking into account that we shall be dealing with graded Hopf algebras, both left and right definitions of Calabi-Yau coincide,
and we will thus omit the reference to the side.
We claim that the image of $(K_{\bullet}(\ym(n,s)^{\Gamma}), b_{\bullet})$
under the functor $\mathcal{H}om_{(\YM(n,s)^{\Gamma})^{e}}(\place, (\YM(n,s)^{\Gamma})^{e})$
is isomorphic to its shift $(K_{\bullet}(\ym(n,s)^{\Gamma}), (-1)^{\bullet} b_{\bullet})[8]$,
in the category of graded $\YM(n,s)^{\Gamma}$-bimodules, which
follows easily as before from the fact that
\[ \mathcal{H}om_{(\YM(n,s)^{\Gamma})^{e}}(\YM(n,s)^{\Gamma} \otimes X \otimes \YM(n,s)^{\Gamma}, (\YM(n,s)^{\Gamma})^{e}) \simeq
\mathcal{H}om_{k}(X, (\YM(n,s)^{\Gamma})^{e}) \simeq \YM(n,s)^{\Gamma} \otimes X^{*} \otimes \YM(n,s)^{\Gamma}, \]
where $X$ is a finite dimensional graded vector space, and an elementary computation with the differential.
We remark that, if $X = \oplus_{i \in \ZZ} X_{i}$, then $X^{*} = \oplus_{i \in \ZZ} (X^{*})_{i}$, where $(X^{*})_{i} = (X_{-i})^{*}$.
This in turn implies that the super Yang-Mills algebra is graded Calabi-Yau of dimension $3$ and of parameter $8$, for $(n,s) \neq (1,0), (1,1)$.
Hence, we have obtained the following:
\begin{proposition}
Let $(n,s) \in (\NN \times \NN_{0}) \setminus \{ (1,0), (1,1) \}$.
The graded algebra $\YM(n,s)^{\Gamma}$ is AS-regular of global dimension $3$ and of Gorenstein parameter $8$.
Moreover, it is also graded Calabi-Yau of the same dimension and of the same parameter.
\end{proposition}
\subsection{Some consequences}
\label{subsec:cons}
\subsubsection{The Hilbert series of the super Yang-Mills algebra}
We shall compute the Hilbert series of both the graded algebra $\YM(n,s)^{\Gamma}$ and the graded Lie algebra $\ym(n,s)^{\Gamma}$.
In order to do so, we first recall that the Hilbert series is an Euler-Poincar\'e map in the category of graded vector spaces (see \cite{Lang02}, Chap. III, \S 8)
and that the Euler-Poincar\'e characteristic of a complex of graded vector spaces, with respect to the Euler-Poincar\'e map given by taking Hilbert series,
coincides with that of its homology (see \cite{Lang02}, Chap. XX, \S 3, Thm. 3.1).
\begin{proposition}
The Hilbert series of the graded algebra $\YM(n,s)^{\Gamma}$ provided with the grading explained in Subsection \ref{subsec:defi}, \textit{i.e.}
such that $x_{i}$ has degree $2$, for all $i = 1, \dots, n$, and $z_{a}$ has degree $3$, for all $a = 1, \dots, s$, is given by
\[ \YM(n,s)^{\Gamma}(t) = \frac{1}{1 - n t^{2} - s t^{3} + s t^{5} + n t^{6} - t^{8}}. \]
\end{proposition}
\noindent\textbf{Proof.}
Since the complex \eqref{eq:comp} of graded vector spaces is exact, its Euler-Poincar\'e characteristic,
with respect to the Euler-Poincar\'e map given by taking Hilbert series, coincides with that of its homology.
Hence,
\[ 1 - \YM(n,s)^{\Gamma}(t) + (n t^{2} + s t^{3}) \YM(n,s)^{\Gamma}(t) - (n t^{6} + s t^{5}) \YM(n,s)^{\Gamma}(t) + t^{8} \YM(n,s)^{\Gamma}(t) = 0, \]
from which the proposition follows.
\qed
We state the following result, which we believe is well-known, but we give its proof because we do not know any specific reference.
\begin{proposition}
Let $V$ be a positively graded vector space with Hilbert series given by $\sum_{j \in \NN} \nu_{j} t^{j}$.
Then,
\[ \nu_{j} = \frac{(-1)^{j}}{j} \sum_{d|j} (-1)^{d} a_{d} \mu(\frac{j}{d}), \]
for all $j \in \NN$, where $\mu$ is the M\"obius function and the coefficients $a_{d}$ are obtained from the formal series
\[ \log(S(V)(t)) = \sum_{d \in \NN} \frac{a_{d}}{d} t^{d}. \]
Equivalently, if $S(V)(t)^{-1}$ is a polynomial of degree $m$ with roots $\lambda_{i}$, $i = 1, \dots, m$,
\[ a_{d} = \sum_{i=1}^{m} \lambda_{i}^{-d}. \]
\end{proposition}
\noindent\textbf{Proof.}
First note that the Hilbert series of the symmetric algebra (in the super sense) of a graded vector space $V$ satisfying $V(t) = \sum_{j \in \NN} \nu_{j} t^{j}$
is given by
\[ S(V)(t) = \frac{\prod_{i \in 2 \NN - 1} (1 + t^{i})^{\nu_{i}}}{\prod_{i \in 2 \NN} (1 - t^{i})^{\nu_{i}}}
= \prod_{i \in \NN} (1 - (-1)^{i} t^{i})^{-(-1)^{i} \nu_{i}}. \]
In consequence, using that $\log(1-t) = - \sum_{i \in \NN} t^{i}/i$, we obtain that
\[ \log(S(V)(t)) = - \sum_{i \in \NN} (-1)^{i} \nu_{i} \log(1 - (-1)^{i} t^{i}) = \sum_{i,j \in \NN} (-1)^{i} \nu_{i} \frac{(-t)^{i.j}}{j}
= \sum_{l \in \NN} \sum_{i|l} (-1)^{i+l} \nu_{i} i \frac{t^{l}}{l}, \]
which means that
\[ a_{d} = (-1)^{d} \sum_{i|d} (-1)^{i} \nu_{i} i. \]
By the M\"obius inversion formula we have thus
\[ \nu_{j} = \frac{(-1)^{j}}{j} \sum_{d|j} (-1)^{d} a_{d} \mu(\frac{j}{d}), \]
as was to be shown.
To prove the last statement, note that, if $S(V)(t)^{-1}$ is a polynomial of the form
\[ S(V)(t)^{-1} = \prod_{i=1}^{m} (1 - \frac{t}{\lambda_{i}}), \]
then
\[ \log(S(V)(t)) = \sum_{i=1}^{m} \sum_{j \in \NN} \frac{1}{j} \Big(\frac{t}{\lambda_{i}}\Big)^{j}, \]
and the proposition is proved.
\qed
As a consequence of the two previous propositions and the PBW theorem for graded Lie algebras we obtain the
\begin{corollary}
\label{cor:hilserym}
Let $\sum_{j \in \NN} \nu(n,s)_{j} t^{j}$ be the Hilbert series of the graded Lie algebra $\ym(n,s)^{\Gamma}$ provided with the grading explained in Subsection \ref{subsec:defi}.
We shall sometimes write $\nu_{j}$ instead of $\nu(n,s)_{j}$.
Then,
\[ \nu_{j} = \frac{(-1)^{j}}{j} \sum_{d|j} (-1)^{d} a_{d} \mu(\frac{j}{d}), \]
for all $j \in \NN$, where the coefficients $a_{d}$ are obtained from the formal series
\[ \log(1 - n t^{2} - s t^{3} + s t^{5} + n t^{6} - t^{8}) = - \sum_{d \in \NN} \frac{a_{d}}{d} t^{d}, \]
or equivalently, if $\lambda_{1}, \dots, \lambda_{8}$ are the roots of $1 - n t^{2} - s t^{3} + s t^{5} + n t^{6} - t^{8}$, we have that
\[ a_{d} = \sum_{i=1}^{8} \lambda_{i}^{-d} = (-1)^{d} \sum_{i=1}^{8} \lambda_{i}^{d}. \]
\end{corollary}
\subsubsection{Some explicit computations of basis elements of a super Yang-Mills algebra}
As a corollary of the previous computation of the Hilbert series of the super Yang-Mills algebra, we will present in this paragraph
some simple calculations of the basis elements of $\ym(3,1)^{\Gamma}$ for low degree homogeneous components.
Note that $\ym(3,1)^{\Gamma}$ cannot be equivariant.
We remark that in this case, under the assumptions explained in Subsection \ref{subsec:defi}, the space of relations can be taken to be of the form
\begin{align*}
r_{0,1} &= [x_{2},[x_{2},x_{1}]] + [x_{3},[x_{3},x_{1}]] - \frac{1}{2} [z_{1},z_{1}],
\\
r_{0,2} &= [x_{1},[x_{1},x_{2}]] + [x_{3},[x_{3},x_{2}]],
\\
r_{0,3} &= [x_{1},[x_{1},x_{3}]] + [x_{2},[x_{2},x_{3}]],
\\
r_{1,1} &= [x_{1},z_{1}].
\end{align*}
We shall first consider the super Yang-Mills algebra $\ym(3,1)^{\Gamma}$ as a graded Lie algebra, and consider the descending sequence of graded ideals $\{ F^{\bullet}\ym(3,1)^{\Gamma} \}_{\bullet \in \NN_{0}}$,
where $F^{j}\ym(3,1)^{\Gamma}$ is the graded vector subspace of $\ym(3,1)^{\Gamma}$ given by the elements of degree greater than or equal to $j+2$.
Our aim is to give explicit bases for the quotients $\ym(3,1)^{\Gamma}/F^{j}\ym(3,1)^{\Gamma}$, for $j = 1, \dots, 7$.
Using Corollary \ref{cor:hilserym} for $\ym(3,1)^{\Gamma}$, the sequence of dimensions $\nu(3,1)_{j}$ ($j \in \NN$) is given by
\[ 0, 3, 1, 3, 2, 6, 6, 12, 15, 33, 42, 77, 114, 213, 314, 555, 876, 1540, 2460, 4242,\dots \]
An ordered basis for the quotient algebra $\ym(3,1)^{\Gamma}/F^{1}\ym(3,1)^{\Gamma}$ of the super Yang-Mills algebra, which is concentrated in degree $2$,
is given by $\B_{1} = \{ x_{1}, x_{2}, x_{3} \}$,
whereas a basis of $\ym(3,1)^{\Gamma}/F^{2}\ym(3,1)^{\Gamma}$ may be defined as $\B_{2} = \{ x_{1}, x_{2}, x_{3}, z_{1} \}$.
Notice that $\ym(3,1)^{\Gamma}/\C^{2}(\ym(3,1)^{\Gamma}) = \ym(3,1)^{\Gamma}/F^{2}\ym(3,1)^{\Gamma}$.
For the quotient $\ym(3,1)^{\Gamma}/F^{3}\ym(3,1)^{\Gamma}$, a possible ordered basis is
\[ \B_{3} = \{ x_{1}, x_{2}, x_{3}, z_{1}, x_{12}, x_{13} , x_{23} \}, \]
where $x_{ij} = [x_{i} , x_{j}]$, ($i, j = 1, 2, 3$).
To prove that it is indeed a basis we must only show that it generates $\ym(3,1)^{\Gamma}/F^{3}\ym(3,1)^{\Gamma}$ (since $\#(\B_{3}) = 7$),
which is obtained using that $x_{ij} = - x_{ji}$.
The quotient $\ym(3,1)^{\Gamma}/F^{4}\ym(3,1)^{\Gamma}$ is a little more interesting, and a possible ordered basis for it is of the form
\[ \B_{4} = \{ x_{1}, x_{2}, x_{3}, z_{1}, x_{12}, x_{13} , x_{23}, y_{2}, y_{3} \}, \]
where we denote $y_{i} = [x_{i},z_{1}]$ ($i = 1, 2, 3$).
Again, it suffices to show that it generates $\ym(3,1)^{\Gamma}/F^{4}\ym(3,1)^{\Gamma}$, for $\#(\B_{4}) = 9$.
But this follows from the fact that $y_{1} = [x_{1},z_{1}] = r_{1,1} = 0$.
Note that $\ym(3,1)^{\Gamma}/\C^{3}(\ym(3,1)^{\Gamma}) = \ym(3,1)^{\Gamma}/F^{4}\ym(3,1)^{\Gamma}$, because the vanishing of $r_{0,1}$ implies that $[z_{1},z_{1}] = 0$.
We claim that the following set is a basis of $\ym(3,1)^{\Gamma}/F^{5}\ym(3,1)^{\Gamma}$:
\[ \B_{5} = \{ x_{1}, x_{2}, x_{3}, z_{1}, x_{12}, x_{13}, x_{23}, y_{2}, y_{3}, x_{112}, x_{221}, x_{113}, x_{123}, x_{312}, [z_{1},z_{1}] \}, \]
where $x_{ijk} = [x_{i},[x_{j},x_{k}]]$.
Indeed, as before, we only have to prove that the previous set it is a system of generators of $\ym(3,1)^{\Gamma}/F^{5}\ym(3,1)^{\Gamma}$, because $\#(\B_{5}) = 15$.
This is direct, and can be seen from the Yang-Mills relations
\[ x_{332} = - x_{112}, \hskip 0.5cm x_{331} = - x_{221} + \frac{1}{2} [z_{1},z_{1}], \hskip 0.5cm x_{223} = - x_{113}, \]
and the relations given by antisymmetry and Jacobi identity, \textit{i.e.} $x_{ijk} = - x_{ikj}$, and $x_{213} = x_{123} + x_{312}$.
Now, a basis of $\ym(3,1)^{\Gamma}/F^{6}\ym(3,1)^{\Gamma}$ can be given by
\[ \B_{6} = \{ x_{1}, x_{2}, x_{3}, z_{1}, x_{12}, x_{13}, x_{23}, y_{2}, y_{3}, x_{112}, x_{221}, x_{113}, x_{123}, x_{312}, [z_{1},z_{1}], y_{12}, y_{13}, y_{22}, y_{23}, y_{32}, y_{33} \}, \]
where we write $y_{ij} = [x_{i},[x_{j},z_{1}]]$.
That this is a basis is immediate, since its cardinality is $21$ and it is a set of generators of $\ym(3,1)^{\Gamma}/F^{6}\ym(3,1)^{\Gamma}$, for $[z_{1},x_{ij}] = y_{ji} - y_{ij}$.
The case of $\ym(3,1)^{\Gamma}/F^{7}\ym(3,1)^{\Gamma}$ is a little more complicated.
We shall prove that
\begin{align*}
\B_{7} = \{ &x_{1}, x_{2}, x_{3}, z_{1}, x_{12}, x_{13} , x_{23}, y_{2}, y_{3}, x_{112}, x_{221}, x_{113}, x_{123}, x_{312}, [z_{1},z_{1}], y_{12}, y_{13}, y_{22}, y_{23}, y_{32}, y_{33}
\\
&x_{1112}, x_{1221}, x_{1113}, x_{1123}, x_{2221}, x_{2113}, x_{2312}, x_{3112}, x_{3221}, x_{3312}, [x_{2},[z_{1},z_{1}]], [x_{3},[z_{1},z_{1}]] \},
\end{align*}
is an ordered basis, where $x_{ijkl} = [x_{i},[x_{j},[x_{k},x_{l}]]]$.
In order to prove that $\B_{7}$ is a basis it suffices again to verify that it is a system of generators of the super vector space
$\ym(3,1)^{\Gamma}/F^{7}\ym(3,1)^{\Gamma}$.
On the one hand, taking into account that
\[ [[x_{i},x_{j}],[x_{k},x_{l}]] = [[[x_{i},x_{j}],x_{k}],x_{l}] + [x_{k},[[x_{i},x_{j}],x_{l}]]
= [x_{l},[x_{k},[x_{i},x_{j}]]] + [x_{k},[x_{l},[x_{j},x_{i}]]]
= x_{lkij} + x_{klji},
\]
that $[z_{1},[x_{i},z_{1}]] = [x_{i},[z_{1},z_{1}]]/2$, and that $[[[x_{i},x_{j}],x_{k}],x_{l}] = [x_{l},[x_{k},[x_{i},x_{j}]]] = x_{lkij}$,
we see that the set $\B_{6} \cup \{ x_{ijkl} : i, j, k, l = 1, 2, 3 \} \cup \{ [x_{2},[z_{1},z_{1}]], [x_{3},[z_{1},z_{1}]] \}$ is a system of generators.
We shall prove that it is generated by $\B_{7}$.
In fact, we only need to prove that the latter generates the set
\[ \{ x_{i112}, x_{i221}, x_{i113}, x_{i123}, x_{i312} : i = 1, 2, 3 \}, \]
because $\{ x_{112}, x_{221}, x_{113}, x_{123}, x_{312} \}$ are generators of the homogeneous elements of degree $6$.
This last statement is direct:
\begin{align*}
x_{3113} &= x_{1221}, \hskip 0.5cm x_{2112} = - x_{1221}, \hskip 0.5cm
x_{2123} = x_{3221} + x_{2312} - x_{1113},
\\
x_{1312} &= \frac{x_{3112} + x_{2113} - x_{1123}}{2}, \hskip 0.95cm x_{3123} = \frac{x_{1112}+x_{2221}-x_{3312}}{2}- \frac{[x_{2},[z_{1},z_{1}]]}{4}.
\end{align*}
\subsubsection{\texorpdfstring{The ideal $\hat{\tym}(n,s)^{\Gamma}$ and some algebraic properties of the super Yang-Mills algebra}{The ideal tym and some algebraic properties of the super Yang-Mills algebra}}
\label{subsec:tym}
In this paragraph we shall obtain an important result which will be useful in the sequel: the graded Lie algebra $\hat{\tym}(n,s)^{\Gamma}$ is free.
We will also compute the Hilbert series of its space of generators.
In order to do so, we shall prove that the homology groups of degree greater than one of the graded Lie algebra $\hat{\tym}(n,s)^{\Gamma}$ with coefficients in $k$ vanish.
At the end, we shall derive several algebraic properties of the super Yang-Mills algebras.
First, we need a subsidiary result, for which we recall that the quotient $\ym(n,s)^{\Gamma}/\hat{\tym}(n,s)^{\Gamma} \simeq V(2)_{0}$ tells us
that $V(2)_{0}$, and hence $S(V(2)_{0})$, is a module over $\YM(n,s)^{\Gamma}$.
\begin{proposition}
\label{prop:homow}
Let $n \geq 2$.
The homology groups $H_{\bullet} (\ym(n,s)^{\Gamma}, S(V(2)_{0}))$ of the graded Lie algebra $\ym(n,s)^{\Gamma}$ with coefficients in
the module $\U(\ym(n,s)^{\Gamma}/\hat{\tym}(n,s)^{\Gamma}) \simeq S(V(2)_{0})$, which is obviously a graded right $\YM(n,s)^{\Gamma}$-module, are given by
\[ H_{\bullet}(\ym(n,s)^{\Gamma},S(V(2)_{0})) \simeq \begin{cases}
k, &\text{if $\bullet = 0$,}
\\
\hat{W}(n,s)^{\Gamma}, &\text{if $\bullet = 1$,}
\\
0, &\text{else,}
\end{cases}
\]
where $\hat{W}(n,s)^{\Gamma}$ is the graded vector space with Hilbert series
\[ \hat{W}(n,s)^{\Gamma}(t) = (n-2) t^{2} + (2n-3) t^{4} + \sum_{k \geq 3} (2n - 4) t^{2 k} + \sum_{k \geq 1} s t^{2 k + 1}. \]
\end{proposition}
\noindent\textbf{Proof.}
We shall make the computation of the homology groups using the minimal projective resolution in Proposition \ref{prop:projres}.
It is direct to see that the complex $(S(V(2)_{0}) \otimes_{\YM(n,s)^{\Gamma}} K_{\bullet}(\ym(n,s)^{\Gamma}), 1 \otimes b_{\bullet}')$ is the direct sum
of two complexes $(K_{\bullet}^{\sim},b_{\bullet}^{\sim})$ and $(K_{\bullet}^{\backsim},b_{\bullet}^{\backsim})$, which we now describe.
The latter is given by $K_{2}^{\backsim} = S(V(2)_{0}) \otimes V(s)^{*}_{1}[-8]$, $K_{1}^{\backsim} = S(V(2)_{0}) \otimes V(s)_{1}$ and zero otherwise,
with differential
\begin{equation}
\label{eq:k''}
S(V(2)_{0}) \otimes V(s)^{*}_{1}[-8] \overset{b_{2}^{\backsim}}{\rightarrow} S(V(2)_{0}) \otimes V(s)_{1}
\end{equation}
of the form
\[ b_{2}^{\backsim} (y \otimes r_{1,a}) = \sum_{i = 1}^{n} \sum_{b = 1}^{s} \Gamma_{a,b}^{i} y x_{i} \otimes z_{b}, \]
where we remind that the action of $x_{i}$ on $y$ is zero for $i = 3, \dots, n$, but we prefer to write the complete sum for convenience.
The complex $(K_{\bullet}^{\sim},b_{\bullet}^{\sim})$ is similar to the one analyzed in \cite{HS10}, Prop. 3.5, (3.3).
It is obvious that $H_{0}(K_{\bullet}^{\sim}) \simeq k$, and further $H_{3}(K_{\bullet}^{\sim})=0$, for the map $b_{3}^{\sim}$ is injective.
Indeed, $b_{3}^{\sim}(y) = 0$, for $y \in S(V(2)_{2})$, means that $y x_{i} = 0$, for $i = 1, 2$, which implies that $y$ vanishes, for $S(V(2)_{0})$ is a domain.
Moreover, we shall also show that $H_{2}(K_{\bullet}^{\sim})=0$.
Consider $y = \sum_{i=1}^{n} y_{i} \otimes r_{0,i} \in K_{2}^{\sim}$ in the kernel of $b_{2}^{\sim}$, which is equivalent to the vanishing of
\[ b_{2}^{\sim}(\sum_{i=1}^{n} y_{i} \otimes r_{0,i}) = \sum_{i,j=1}^{n} y_{i} x_{j}^{2} \otimes x_{i} - y_{i} x_{j} x_{i} \otimes x_{j}
= \sum_{i,j=1}^{n} (y_{i} x_{j}^{2} - y_{j} x_{i} x_{j}) \otimes x_{i}. \]
This means that $\sum_{j=1}^{n} (y_{i} x_{j}^{2} - y_{j} x_{i} x_{j}) = 0$, for all $i = 1, \dots, n$.
Since the action of $x_{i}$ on $S(V(2)_{0})$ is zero for $i = 3, \dots, n$, we get that $y_{i} (\sum_{j=1}^{2} x_{j}^{2}) = 0$, for all $i = 3, \dots, n$,
which in turn implies that $y_{i}$ vanishes for $i = 3, \dots, n$.
Hence, the cycle $y$ has in fact the form $\sum_{i=1}^{2} y_{i} \otimes r_{0,i}$, and satisfies that
$\sum_{j=1}^{2} (y_{i} x_{j}^{2} - y_{j} x_{i} x_{j}) = 0$, for $i = 1, 2$.
So, $y$ can be regarded as a cycle of the complex analyzed in \cite{HS10}, Prop. 3.5, (3.3), for $n=2$, whose second homology group
vanishes.
By the definition of the differential of this latter complex, we conclude that there exists $x \in S(V(2)_{0})$ such that $y_{i} = x x_{i}$, for $i = 1, 2$,
which tells us that $b_{3}^{\sim}(x) = y$.
We shall further prove that $H_{\bullet}(K_{\bullet}^{\backsim})$ also vanishes for $\bullet = 2$, and so for $\bullet \geq 2$.
Otherwise stated, we will prove that $b_{2}^{\backsim}$ is injective.
Since $b_{2}^{\backsim}$ is a morphism between two finitely generated free modules (of the same rank) over the commutative domain $S(V_{2})$,
$b_{2}^{\backsim}$ is injective if and only if its determinant $\det(b_{2}^{\backsim})$ does not vanish.
Since the matrix representation of $b_{2}^{\backsim}$ in the respective bases $\{r_{1,a}\}_{a = 1, \dots, s}$ and $\{ z_{a} \}_{a = 1, \dots, s}$ is given by
$(x_{1} \delta_{a,b} + \Gamma_{a,b}^{2} x_{2})_{a,b}$, its determinant has a term of the form $x_{1}^{s}$, and hence it does not vanish.
As a consequence, $H_{\bullet}(K_{\bullet}^{\backsim})$ vanishes for $\bullet = 2$, and thus $H_{\bullet}(\ym(n,s)^{\Gamma},S(V(2)_{0})) = 0$, for $\bullet \geq 2$.
It remains to prove that the Hilbert series of $\hat{W}(n,s)^{\Gamma} = H_{1}(\ym(n,s)^{\Gamma},S(V(2)_{0}))$ is the stated above.
By the previous remarks, we know that $H_{1}(\ym(n,s)^{\Gamma},S(V(2)_{0})) \simeq H_{1}(K_{\bullet}^{\sim}) \oplus H_{1}(K_{\bullet}^{\backsim})$.
This follows from the fact that the Euler-Poincar\'e characteristic of the complex
$(S(V(2)_{0}) \otimes_{\YM(n,s)^{\Gamma}} K_{\bullet}(\ym(n,s)^{\Gamma}), 1 \otimes b_{\bullet}')$ of graded vector spaces,
with respect to the Euler-Poincar\'e map given by taking Hilbert series,
coincides with that of its homology, \textit{i.e.}
\[ 1 - \hat{W}(n,s)^{\Gamma}(t) = \frac{1 - n t^{2} - s t^{3} + n t^{6} + s t^{5} - t^{8}}{(1-t^{2})^{2}}. \]
The Hilbert series of $\hat{W}(n,s)^{\Gamma}$ stated at the beginning follows easily from the previous identity, and the proposition is thus proved.
\qed
\begin{remark}
If the super Yang-Mills is equivariant, the injectivity of $b_{2}^{\backsim}$ can also be proved as follows.
Consider the map
\begin{align*}
S(V(n)_{0}) \otimes V(s)_{1} &\overset{(b_{2}^{\backsim})^{*}}{\rightarrow} S(V(n)_{0}) \otimes V(s)^{*}_{1}[-8]
\\
z \otimes z_{a} &\mapsto \sum_{i = 1}^{n} \sum_{b = 1}^{s} \tilde{\Gamma}^{i,a,b} z x_{i} \otimes r_{1,b},
\end{align*}
where we write again all the variables in the previous sum for convenience, even though the action of $x_{i}$ on $S(V(2)_{0})$ vanishes for $i = 3, \dots, n$.
Condition \eqref{eq:relgammasind} tells us that $((b_{2}^{\backsim})^{*} \circ b_{2}^{\backsim})(z \otimes r_{1,a}) = z.(\sum_{i=1}^{n} x_{i}^{2}) \otimes r_{1,a}$,
which in turn implies that $b_{2}^{\backsim}$ is injective, for $z.(\sum_{i=1}^{n} x_{i}^{2}) = z.(\sum_{i=1}^{2} x_{i}^{2})$ and
$S(V(2)_{0})$ is a domain.
\end{remark}
As a consequence of the previous proposition we obtain the following.
\begin{theorem}
\label{teo:libre}
Let $n \geq 2$.
The graded Lie algebra $\hat{\tym}(n,s)^{\Gamma}$ is free with space of generators isomorphic to $\hat{W}(n,s)^{\Gamma}$.
\end{theorem}
\noindent\textbf{Proof.}
Since $\U(\ym(n,s)^{\Gamma}) \otimes_{\U(\hat{\tym}(n,s)^{\Gamma})} k \simeq S(V(2)_{0})$, Schapiro's Lemma for graded Lie algebras
(which is proved in the same way as for Lie algebras or group algebras, \textit{cf.} \cite{Wei94}, Lemma 6.3.2)
implies that $H_{\bullet} (\ym(n,s)^{\Gamma},S(V(2)_{0})) \simeq H_{\bullet}(\hat{\tym}(n,s)^{\Gamma},k)$.
The previous computation tells us that $H_{\bullet}(\hat{\tym}(n,s)^{\Gamma},k) \simeq \Tor^{\hat{\TYM}(n,s)^{\Gamma}}_{\bullet} (k,k) = 0$, for $\bullet \geq 2$,
so the minimal projective resolution of the $\hat{\TYM}(n,s)^{\Gamma}$-module $k$, which is of the form
$(\Tor_{\bullet}^{\hat{\TYM}(n,s)^{\Gamma}}(k,k),d_{\bullet})$,
has only two nonvanishing components, for $\bullet = 0, 1$.
Hence $H_{\bullet}(\hat{\tym}(n,s)^{\Gamma},M) = 0$, for $\bullet \geq 2$ and for any $\hat{\tym}(n,s)^{\Gamma}$-module $M$,
which implies that $\hat{\tym}(n,s)^{\Gamma}$ is a free graded Lie algebra
(this is proved in the same way as for Lie algebras, \textit{cf.} \cite{Wei94}, Exer. 7.6.3).
Moreover, Shapiro's Lemma tells us that $\hat{\tym}(n,s)^{\Gamma} = \f(\hat{W}(n,s)^{\Gamma})$, for
$\hat{W}(n,s)^{\Gamma} = \hat{\tym}(n,s)^{\Gamma}/[\hat{\tym}(n,s)^{\Gamma},\hat{\tym}(n,s)^{\Gamma}] \simeq H_{1}(\hat{\tym}(n,s)^{\Gamma},k)
\simeq H_{1}(\ym(n,s)^{\Gamma}, S(V(2)_{0}))$ and the theorem follows.
\qed
\begin{remark}
Since the Lie ideal $\tym(n,s)^{\Gamma}$ is a subalgebra of the free graded Lie algebra $\hat{\tym}(n,s)^{\Gamma}$, it is also a free graded Lie algebra,
by Shapiro's Lemma.
As in the proof of the theorem, taking into account that $\U(\ym(n,s)^{\Gamma}) \otimes_{\U(\tym(n,s)^{\Gamma})} k \simeq S(V(n)_{0})$,
Shapiro's Lemma tells us that $H_{\bullet} (\ym(n,s)^{\Gamma},S(V(n)_{0})) \simeq H_{\bullet}(\tym(n,s)^{\Gamma},k)$.
Since $\tym(n,s)^{\Gamma}$ is free, the latter homology group should vanish for $\bullet \geq 2$, $H_{0}(\tym(n,s)^{\Gamma},k) \simeq k$, and
$H_{1}(\tym(n,s)^{\Gamma},k) \simeq H_{1}(\ym(n,s)^{\Gamma},S(V(n)_{0})) \simeq W(n,s)^{\Gamma}$,
where $W(n,s)^{\Gamma}$ is the graded vector space of generators of $\tym(n,s)^{\Gamma}$.
Its Hilbert series can be computed using again the fact that the Euler-Poincar\'e characteristic of a complex of graded vector spaces,
with respect to the Euler-Poincar\'e map given by taking Hilbert series, coincides with that of its homology, applied to the complex
$(S(V(n)_{0}) \otimes_{\YM(n,s)^{\Gamma}} K_{\bullet}(\ym(n,s)^{\Gamma}), 1 \otimes b_{\bullet}')$.
We get that
\[ W(n,s)^{\Gamma}(t) = \frac{(1-t^{2})^{n}-1+n t^{2} + s t^{3} - s t^{5} - n t^{6} +t^{8}}{(1-t^{2})^{n}}. \]
Note that the graded vector space given by the even part of $W(n,s)^{\Gamma}$ is isomorphic to the graded vector space $W(n)$ considered in \cite{HS10h}, Section 3,
provided with the special grading.
\end{remark}
For completeness, we shall also analyze the case of super Yang-Mills algebras $\ym(n,s)^{\Gamma}$ when $n=1$.
In the case $s = 0$, the super Yang-Mills algebra $\ym(1,0)^{0}$ is just the one-dimensional super Lie algebra concentrated in degree zero.
We shall now restrict ourselves to the case $s \neq 0$.
As noted before, the nondegeneracy of $\Gamma$ implies that $\ym(1,s)^{\Gamma} \simeq k.x_{1} \times \h(1,s)$,
where $\h(1,s) \simeq \f(z_{1},\dots,z_{s})/\cl{\sum_{a=1}^{s} [z_{a},z_{a}]}$.
The cases $\ym(1,1)$ and $\ym(1,2)$ are nilpotent and finite dimensional, and its representation theory can be analyzed using the below recalled
Kirillov orbit method.
In particular, $\ym(1,1)^{\Gamma}$ is a supercommutative super Lie algebra of super dimension $(1,1)$.
The super Lie algebra $\ym(1,2)^{\Gamma}$ has super dimension $(2,2)$, with basis $x_{1}$, $z_{1}$, $z_{2}$ and $[z_{2},z_{2}]$, where $[z_{1},z_{1}]=-[z_{2},z_{2}]$, and all other brackets vanish.
As a consequence, since the enveloping algebra of a finite dimensional super Lie algebra is noetherian, we get that $\YM(1,1)^{\Gamma}$
and $\YM(1,2)^{\Gamma}$ are noetherian.
We shall now suppose that $s \geq 3$.
In this case, it is easy to prove that the Lie ideal $\k(1,n)$ of $\h(1,s)$ given by $\cl{z_{3}, \dots, z_{s}, [z_{1},z_{2}]} + F^{7}\h(1,n)$, where
$F^{7}\h(1,n)$ is the super vector space formed of elements of $\h(1,s)$ of degree greater than or equal to $9$.
It is also an ideal when regarded inside $\ym(1,s)^{\Gamma}$.
Note that $\U(\h(1,n)/\k(1,n)) \simeq k\cl{z_{1}, z_{2}}/\cl{z_{1}^{2}+z_{2}^{2},[z_{1},z_{2}]}$, where the $[z_{1},z_{2}]$ is the supercommutator of
$z_{1}$ and $z_{2}$.
Using the Diamond Lemma, it is easy to see that a (homogeneous) basis of it is given by $z_{1}^{\alpha} z_{2}^{\beta}$,
where $\alpha \in \{ 0 , 1\}$, and $\beta \in \NN_{0}$, and the multiplication is determined by
\[ z_{2}^{\beta} z_{1} z_{2}^{\beta'} = (-1)^{\beta} z_{1} z_{2}^{\beta + \beta'}, \]
and
\[ z_{1} z_{2}^{\beta} z_{1} z_{2}^{\beta'} = (-1)^{\beta} z_{2}^{\beta + \beta' + 2}. \]
The following proposition is a direct consequence of the expression of the complex \eqref{eq:hns} and the previous description of $\U(\h(1,n)/\k(1,n))$.
\begin{proposition}
Let $s \geq 3$.
The homology groups $H_{\bullet} (\h(1,s), \U(\h(1,n)/\k(1,s)))$ of the graded Lie algebra $\h(1,s)$ with coefficients in
the module $\U(\h(1,n)/\k(1,n))$, which is obviously a graded right $H(1,s)$-module, are given by
\[ H_{\bullet} (\h(1,s), \U(\h(1,s)/\k(1,s))) \simeq \begin{cases}
k, &\text{if $\bullet = 0$,}
\\
\tilde{W}(1,s), &\text{if $\bullet = 1$,}
\\
0, &\text{else,}
\end{cases}
\]
where $\tilde{W}(1,s)$ is the graded vector space with Hilbert series
\[ \tilde{W}(1,s)(t) = (s-2) t^{3} + (2s-3) t^{6} + \sum_{k \geq 3} (2s - 4) t^{3 k}. \]
\end{proposition}
The arguments used in Theorem \ref{teo:libre} also yield
\begin{theorem}
\label{teo:libre2}
Let $s \geq 3$.
The subalgebra $\k(1,s)$ of the graded Lie algebra $\h(1,n)$, and also of $\ym(1,s)^{\Gamma}$, is free with space of generators isomorphic to
$\tilde{W}(1,s)$.
\end{theorem}
As a corollary of the previous theorems we obtain the following result.
\begin{corollary}
\label{cor:nonoeth}
Let $n \geq 2$ or $n = 1$ and $s \geq 3$.
In either case the super algebra $\YM(n,s)^{\Gamma}$ is not (left nor right) noetherian.
\end{corollary}
\noindent\textbf{Proof.}
Let us denote by $\h$ a free super Lie algebra inside $\ym(n,s)^{\Gamma}$ such that the dimension of (the underlying vector space of) $\h$ is greater than $1$,
which exists by the previous theorems.
The proposition is now proved by an analogous argument as the one presented in the three paragraphs of \cite{HS10} before Remark 3.14.
Since the extension $\U(\ym(n,s)^{\Gamma}) \supseteq \U(\h)$ is free
(\textit{i.e.} $\U(\ym(n,s)^{\Gamma})$ is a free (left) module over $\U(\h)$), any (left) ideal $I \subseteq \U(\h)$ satisfies that
$\U(\ym(n,s)^{\Gamma}).I \cap \U(\h) = I$.
The fact that the super algebra $\U(\h)$ is not noetherian (because it is free with more than one generator) implies that the super algebra $\U(\ym(n,s)^{\Gamma})$ is not noetherian.
\qed
Even though the previous algebras are not noetherian, we will prove that they are (left and right) coherent.
We recall that a (say left) finitely generated module $M$ over graded algebra $A$ is called \emph{coherent} if all its finitely generated submodules are finitely presented.
It is easy to see that the category of coherent modules over a graded ring is an abelian subcategory of the category of all modules ${}_{A}\Mod$
over the graded algebra $A$.
Moreover, the graded algebra $A$ is called left (resp. right) \emph{coherent} if the category of finitely presented left (resp. right) modules coincides with the category of coherent modules, or otherwise stated if the former is abelian.
We stress that all modules here are graded.
Also note that the previous condition can be easily seen to be equivalent to either of the following statements:
any finitely generated submodule of a finitely presented left (resp. right) module is finitely presented, or any finitely generated left (resp. graded) ideal is finitely presented.
We shall denote by ${}_{A}\mod$ the category of coherent (graded) modules over the graded algebra $A$.
Note that the super Yang-Mills algebras are (graded) Hopf algebras, for which the notions of left and right noetherianity coincide,
and the same applies to coherency.
So, we shall just refer to these algebras as noetherian, or coherent.
We first recall the following result due to D. Piontkovski.
\begin{proposition}
\label{prop:piont}
Let $A$ be a graded algebra and $J$ a two-sided ideal of $A$ (different from $A$), which is free as a left module.
Then, if the quotient graded algebra $A/J$ is right noetherian, $A$ is right coherent.
\end{proposition}
\noindent\textbf{Proof.}
See \cite{Pi08}, Prop. 3.2.
\qed
We have the following immediate consequence.
\begin{corollary}
\label{coro:coh}
Let $\g$ be a $\NN$-graded Lie algebra and $\h$ a Lie ideal of $\g$ ($\h \neq \g$).
Then, the $\U(\g)$-module $\U(\g)\h$ is free if and only if $\h$ is a free graded Lie algebra.
As a consequence, assuming that the quotient graded algebra $\U(\g/\h)$ is noetherian and $\h$ is a free graded Lie algebra, $\U(\g)$ is coherent.
\end{corollary}
\noindent\textbf{Proof.}
Note first that $\U(\g)\h = \h \U(\g)$ is a two-sided ideal, and $\U(\g/\h) \simeq \U(\g)/(\U(\g)\h)$, so the second statement follows directly from the first one and Proposition \ref{prop:piont}.
In consequence, we only have to prove that $\U(\g)\h$ is a free $\U(\g)$-module if and only if $\h$ is a free graded Lie algebra.
The former is equivalent to show that $\Tor^{\U(\g)}_{\bullet}(k,\U(\g)\h)$ vanishes for $\bullet \geq 1$, since $\U(\g) \h$ is a bounded below graded $\U(\g)$-module,
and the latter is equivalent to the vanishing of $\Tor^{\U(\h)}_{\bullet}(k,k) \simeq \Tor^{\U(\g)}_{\bullet}(k,\U(\g/\h))$, for $\bullet \geq 2$, since $\U(\h)$ is a graded connected algebra,
as explained in Theorem \ref{teo:libre}.
The previous isomorphism follows from Schapiro's Lemma, since
$\Tor^{\U(\h)}_{\bullet}(k,k) \simeq H_{\bullet}(\h,k) \simeq H_{\bullet}(\g,\U(\g/\h)) \simeq \Tor^{\U(\g)}_{\bullet}(k,\U(\g/\h))$.
It thus suffices to prove that both homology groups $\Tor^{\U(\g)}_{\bullet}(k,\U(\g)\h)$ and $\Tor^{\U(\g)}_{\bullet+1}(k,\U(\g/\h))$ are isomorphic, for $\bullet \geq 1$.
Consider the short exact sequence of $\U(\g)$-modules
\[ 0 \rightarrow \U(\g) \h \rightarrow \U(\g) \rightarrow \U(\g/\h) \rightarrow 0, \]
which induces a long exact sequence on torsion groups
\[ \dots \rightarrow \Tor_{i+1}^{\U(\g)}(k,\U(\g/\h)) \rightarrow \Tor_{i}^{\U(\g)}(k,\U(\g)\h) \rightarrow \Tor_{i}^{\U(\g)}(k,\U(\g)) \rightarrow \Tor_{i}^{\U(\g)}(k,\U(\g/\h)) \rightarrow \Tor_{i-1}^{\U(\g)}(k,\U(\g)\h) \rightarrow \dots \]
The vanishing of $\Tor_{\bullet}^{\U(\g)}(k,\U(\g))$ for $\bullet \geq 1$ (because $\U(\g)$ is free) implies that
$\Tor_{\bullet+1}^{\U(\g)}(k,\U(\g/\h)) \simeq \Tor_{\bullet}^{\U(\g)}(k,\U(\g)\h))$ for $\bullet \geq 1$,
which proves the corollary.
\qed
Theorems \ref{teo:libre} and \ref{teo:libre2} together with the previous corollary now yield the result:
\begin{corollary}
Let $n \geq 2$, or $n = 1$ and $s \geq 3$.
Then the super Yang-Mills algebra $\YM(n,s)^{\Gamma}$ is (left and right) coherent.
\end{corollary}
\begin{remark}
\label{rem:coh}
Note that $\ym(0,s)^{0}$ is a free graded algebra on the generators $z_{1}, \dots, z_{s}$, and hence it is obviously seen to be coherent (applying the Corollary \ref{coro:coh} for $\h = \f(z_{2}, \dots, z_{s})$).
For the other values of the parameters ($n = 1$ and $s = 1$ or $s = 2$), the super Yang-Mills algebra is noetherian, so \textit{a fortiori} coherent.
As a consequence, we see that the super Yang-Mills algebras are coherent for all values of the parameters $n$ and $s$.
\end{remark}
Since $\YM(n,s)^{\Gamma}$ is AS-regular with Gorenstein parameter $8$,
we may consider the (finite dimensional) \emph{Beilinson algebra} $\nabla \YM(n,s)^{\Gamma}$ associated to it,
\textit{i.e.}
\[ \nabla \YM(n,s)^{\Gamma} = \begin{pmatrix}
\YM(n,s)^{\Gamma}_{0} & \YM(n,s)^{\Gamma}_{1} & \dots & \YM(n,s)^{\Gamma}_{7}
\\
0 & \YM(n,s)^{\Gamma}_{0} & \dots & \YM(n,s)^{\Gamma}_{6}
\\
\vdots & \vdots & \dots & \vdots
\\
0 & 0 & \dots & \YM(n,s)^{\Gamma}_{0}
\end{pmatrix}, \]
with the obvious matrix multiplication.
Equivalently, it can also be easily defined by a quiver algebra with relations.
We now recall for a connected $\NN_{0}$-graded algebra $A$ the definition of the (abelian) quotient categories
$\mathrm{Tails}(A) = {}_{A}\Mod/{}_{A}\mathrm{Tors}$ and $\mathrm{tails}(A) = {}_{A}\mod/{}_{A}\mathrm{tors}$,
where ${}_{A}\mathrm{Tors}$ is the category of torsion (graded) modules over $A$,
\textit{i.e.} the modules $M$ that satisfy that for any $m \in M$, there is $i \in \NN$ such that $A_{\geq i}.m = 0$,
and ${}_{A}\mathrm{tors}$ is the category of torsion coherent modules over $A$ (\textit{cf.} \cite{AZ94}, Section 1).
Then, using \cite{MM11}, Thm. 4.12 and 4.14, we obtain the
\begin{proposition}
Let $n \geq 2$, or $n = 1$ and $s \geq 3$.
Then, there exists equivalences of triangulated categories
\begin{align*}
D(\mathrm{Tails}(\YM(n,s)^{\Gamma})) &\simeq D({}_{\nabla \YM(n,s)^{\Gamma}}\Mod),
\\
D^{b}(\mathrm{tails}(\YM(n,s)^{\Gamma})) &\simeq D^{b}({}_{\nabla \YM(n,s)^{\Gamma}}\mod).
\end{align*}
\end{proposition}
As a further consequence of the freeness of the subalgebra $\hat{\tym}(n,s)^{\Gamma}$, we obtain the following result.
\begin{proposition}
Let $n \geq 2$.
Then, the super Yang-Mills algebra $\YM(n,s)^{\Gamma}$ is a semiprimitive domain, \textit{i.e.}, it has vanishing Jacobson radical and does not have zero divisors.
\end{proposition}
\noindent\textbf{Proof.}
We first prove that the super Yang-Mills algebra is a domain.
In order to do so, note that, since all odd elements of the super Yang-Mills algebra $\ym(n,s)^{\Gamma}$ are in fact included in the free super Lie algebra $\hat{\tym}(n,s)^{\Gamma}$,
we have that $[y,y] \neq 0$, for all its odd elements $y$.
Then, \cite{AL85}, Thm. 2.7, implies that $\YM(n,s)^{\Gamma}$ is a domain.
We shall now prove that the super Yang-Mills algebra is semiprimitive, \textit{i.e.} the Jacobson radical of $\YM(n,s)^{\Gamma}$ vanishes.
We recall that the Jacobson radical of a super algebra and of its underlying algebra coincide (see \cite{CM84}, Thm. 4.4, (3)).
Also note that the Jacobson radical of a free algebra is zero, since for any nonzero element $w$ of a free algebra, $1 - u w$ cannot be invertible for all elements $u$ in the free algebra
(take $u$ to be nonzero and noninvertible).
Let $n \geq 2$.
Then the collection of one-codimensional inclusions of Lie ideals
$\hat{\tym}(n,s)^{\Gamma} \subseteq \h(n,s) \subseteq \ym(n,s)^{\Gamma}$ tells US that $\U(\h(n,s)) \simeq \U(\hat{\tym}(n,s)^{\Gamma})[x_{2},\delta_{2}]$,
where $\delta_{2}$ is the derivation on $\U(\hat{\tym}(n,s)^{\Gamma})$ induced by $\ad(x_{2})$, and
$\U(\ym(n,s)^{\Gamma}) \simeq \U(\h(n,s))[x_{1},\delta_{1}]$, where $\delta_{1}$ is the derivation on $\U(\h(n,s))$ induced by $\ad(x_{1})$.
As a consequence, the super Yang-Mills algebra $\U(\ym(n,s)^{\Gamma})$ is a sequence of Ore extensions with derivations of a free algebra, for $n \geq 2$.
Also, note that, if $A \subseteq B$, is an Ore extension, then it is free, and in particular, by the proof of Corollary \ref{cor:nonoeth}, for
any left ideal $I$ of $A$, we have that $B.I \cap A = I$.
This in turn implies that $J(B) \cap A \subseteq J(A)$.
Indeed, consider the collection $\mathcal{S}$ of maximal left ideals $J$ of $B$ such that they contain $B.I$, for some $I$ a maximal left ideal of $A$.
Note that $B.I \neq B$, since $B.I \cap A = I$.
Moreover, if $J \in \mathcal{S}$, and $J \supseteq B.I$, for $I$ a maximal ideal of $A$, then $I = B.I \cap A \subseteq J \cap A \neq A$, and hence
$J \cap A = I$, for $I$ is maximal.
We remark that $J \cap A \neq A$, because $1_{A} = 1_{B} \notin J$.
Let $\mathcal{S}'$ denote the collection of all maximal left ideals of $A$.
Since $J(B)$ is the intersection of all maximal left ideals of $B$, then
\[ J(B) \cap A \subseteq \bigcap_{J \in \mathcal{S}} J \cap A = \bigcap_{I \in \mathcal{S}'} I = J(A). \]
The proposition now follows from \cite{FKM83}, Thm. 3.2.
\qed
\section{The main result on representations of super Yang-Mills algebra}
\label{sec:main}
The aim of this last section is to prove that (most of) the Clifford-Weyl super algebras $\Cliff_{q}(k) \otimes A_{p}(k)$
are epimorphic images of all super Yang-Mills algebras $\YM(n,s)^{\Gamma}$, under certain assumptions.
This will rely on our previous study of the Lie ideal $\hat{\tym}(n,s)^{\Gamma}$.
As a consequence, the representations of such Clifford-Weyl super algebras $\Cliff_{q}(k) \otimes A_{p}(k)$
are also representations of $\YM(n,s)^{\Gamma}$, which is the analogous result to the one proved in \cite{HS10}.
On the one hand, since the super algebra $\Cliff_{2q}(k) \simeq M_{2^{q}}(k)$ is Morita equivalent to the super algebra $k$,
we easily conclude that the Clifford-Weyl super algebra $\Cliff_{2q}(k) \otimes A_{p}(k)$ is Morita equivalent to the super algebra $A_{p}(k)$, and
$\Cliff_{2q + 1}(k) \otimes A_{p}(k)$ is Morita equivalent to $\Cliff_{1}(k) \otimes A_{p}(k)$.
Furthermore, the category of representations of the super algebra $\Cliff_{1}(k) \otimes A_{p}(k)$ is equivalent to the category of representations of the Weyl algebra $A_{p}(k)$.
In both cases we see that the representations we obtain can be understood as induced by those of the Weyl algebra (see Remark \ref{rem:morita}).
\subsection{Some prerequisites}
We shall now briefly recall a version of the Kirillov orbit method and the Dixmier map for nilpotent super Lie algebras, which we will employ.
We shall use the conventions of \cite{Her10}, to which we refer for the details and the bibliography therein.
We recall that a bilateral ideal $I \triangleleft A$ of a super algebra $A$ is called \emph{primitive} if it is the annihilating ideal of a simple $A$-module,
and it is called \emph{maximal} if $I \neq A$ and if it is maximal in the lattice of bilateral ideals of $A$, ordered by inclusion.
Every maximal ideal is clearly primitive.
We have the following proposition.
\begin{proposition}
\label{prop:lieweyl}
Let $I \triangleleft \U(\g)$ be a bilateral ideal of the universal enveloping algebra of a nilpotent super Lie algebra $\g$ of finite dimension.
The following are equivalent:
\begin{itemize}
\item[(i)] $I$ is primitive.
\item[(ii)] $I$ is maximal.
\item[(iii)] There exist $p, q \in \NN_{0}$ such that $\U(\g)/I \simeq \Cliff_{p}(k) \otimes A_{q}(k)$, where $\Cliff_{p}(k)$ is the Clifford (super) algebra over $k$
and $A_{q}(k)$ is the $q$-th Weyl algebra, which is concentrated in degree zero.
\item[(iv)] $I$ is the kernel of a simple representation of $\U(\g)$.
\end{itemize}
\end{proposition}
\noindent\textbf{Proof.}
See \cite{Let89}, Prop. 3.3, and \cite{Her10}, Prop. 4.13.
\qed
We remark that the Clifford-Weyl super algebra $\Cliff_{q}(k) \otimes A_{p}(k)$ in the previous theorem has the $\ZZ/2\ZZ$-grading induced by usual grading of the Clifford (super) algebra
$\Cliff_{q}(k)$ and by considering the Weyl algebra $A_{p}(k)$ to be concentrated in degree zero.
If $I \triangleleft \U(\g)$ is a bilateral ideal satisfying either of the previous equivalent conditions, the pair of nonnegative integers $(p,q)$
(uniquely determined) such that $\U(\g)/I \simeq \Cliff_{p}(k) \otimes A_{q}(k)$ is called the \emph{weight} of the ideal $I$.
Let us suppose that $\g$ is a nilpotent super Lie algebra of finite dimension.
Given $f \in \g^{*}_{0}$, a \emph{polarization} of $\g$ at $f$ is a subalgebra $\h \subset \g$ such that it is \emph{subordinated} to $f$,
\textit{i.e.} $f([\h,\h]) = 0$,
and it is in fact a maximal subspace of the super vector space underlying $\g$ with respect to the property that
the bilinear form $f([\place,\place])$ vanishes on it.
It is easily verified that the super dimensions of all polarizations of $\g$ at $f$ coincide and, furthermore,
that $\h$ is a polarization of $\g$ at $f$ if and only if $\h$ is a subalgebra subordinated to $f$ whose super dimension coincides
with the one of a polarization of $\g$ at $f$ (see \cite{Her10}, Subsection 3.4).
We shall now explain the connection between primitive ideals and even linear functionals for a nilpotent super Lie algebra.
If $f \in \g^{*}_{0}$ is a linear functional and $\h_{f}$ a polarization at $f$,
we may define a representation of $\h_{f}$ on the vector space $k.v_{f}$ of dimension $1$ by means of
$x.v_{f} = f(x) v_{f}$, for $x \in \h_{f}$.
Therefore, we can consider the induced $\U(\g)$-module $V_{f} = \U(\g) \otimes_{\U(\h_{f})} k.v_{f}$.
If we denote the corresponding action by $\rho : \U(\g) \rightarrow \End_{k}(V_{f})$, $I(f) = \Ker(\rho)$ is a bilateral ideal of the
enveloping algebra $\U(\g)$.
In the previous notation we have omitted the polarization in $I(f)$, due to the following proposition, which states even more.
\begin{proposition}
Let $\g$ be a nilpotent super Lie algebra of finite dimension, let $f \in \g^{*}_{0}$ and let $\h_{f}$ and $\h'_{f}$ be two polarizations of $f$.
If we denote $\rho : \U(\g) \rightarrow \End_{k}(V_{f})$ and $\rho' : \U(\g) \rightarrow \End_{k}(V'_{f})$
the corresponding representations constructed following the previous method, then $\Ker(\rho) = \Ker(\rho')$.
This ideal is primitive.
On the other hand, if $I$ is a primitive ideal of $\U(\g)$, then there exists $f \in \g^{*}_{0}$ such that $I = I(f)$.
\end{proposition}
\noindent\textbf{Proof.}
See \cite{Her10}, Thm. 4.5, Thm. 4.7 and Thm. 4.9.
\qed
The weight of a primitive ideal $I(f)$ is given by $(p,q) = (\dim(\g_{0}/\g^{f}_{0})/2, \dim(\g_{1}/\g^{f}_{1}))$,
where $\g^{f} = (\g^{f}_{0}, \g^{f}_{1})$ is the kernel of the superantisymmetric bilinear form $f([\place,\place])$ determined by $f$ on $\g$
(see \cite{Her10}, Prop. 4.13).
The group $\Aut(\g_{0})$ is an algebraic group whose associated Lie algebra is $\Der(\g_{0})$.
Since the super Lie algebra $\g$ is nilpotent, the Lie algebra given by its even part $\g_{0}$ is also so, and Lie algebra given by the ideal of inner derivations $\InnDer(\g_{0})$ in $\Der(\g_{0})$ is algebraic.
The irreducible algebraic group $\mathcal{A}d_{0}$ associated to $\InnDer(\g_{0})$ is called the \emph{adjoint (algebraic) group} of $\g_{0}$.
It is a subgroup of $\Aut(\g_{0})$.
As a consequence, the group $\mathcal{A}d_{0}$ acts on the Lie algebra $\g_{0}$, so it also acts on $\g^{*}_{0}$ with the dual action, which is called \emph{coadjoint}.
\begin{proposition}
Let $\g$ be a nilpotent super Lie algebra of finite dimension and let $f$ and $f'$ be two even linear forms on $\g$, \textit{i.e.} $f, f' \in \g^{*}_{0}$.
If $I(f)$ and $I(f')$ are the corresponding bilateral ideals of $\U(\g)$, then $I(f) = I(f')$ if and only if there is $g \in \mathcal{A}d_{0}$ such that
$f = g.f'$.
\end{proposition}
\noindent\textbf{Proof.}
See \cite{Her10}, Prop. 4.12.
\qed
The previous results imply that, for a nilpotent super Lie algebra of finite dimension there exists an explicit bijection
\[ I : \g^{*}_{0}/\mathcal{A}d_{0} \rightarrow \Prim (\U(\g)) \]
between the set of equivalence classes of even linear forms on $\g$ under the coadjoint action and the set of primitive ideals of the super algebra $\U(\g)$.
We also recall that, if $\g$ be a finite dimensional nilpotent super Lie algebra and $\h$ a Lie ideal,
given $I \triangleleft \U(\h) \subset \U(\g)$ a two sided ideal in the enveloping algebra of $\h$, one defines the \emph{stabilizer} of the ideal $I$ in $\g$ to be
\[ \mathfrak{st}(I, \g) = \{ x \in \g : [x , I] \subset I \}. \]
Let us further suppose that there exists $f \in \g_{0}^{*}$ such that $I = I(f|_{\h_{0}})$.
By \cite{Her10}, Prop. 4.16, if $\g' = \{ x \in \g : f([x,\h])=0 \}$, then
\[ \mathfrak{st}(I, \g) \supseteq \g' + \h. \]
\subsection{Main theorem}
We have first the following result:
\begin{proposition}
\label{teo:yangmillsweyl}
Let $n, s, p \in \NN$ be positive integers, satisfying that $n \geq 3$.
There exists a surjective homomorphism of super algebras
\[ \U(\ym(n,s)^{\Gamma}) \twoheadrightarrow A_{p}(k). \]
Furthermore, there exists $l \in \NN$ such that we can choose this homomorphism satisfying that it factors through the quotient
$\U(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$
\[
\xymatrix@C-20pt@R-10pt { \U(\ym(n,s)^{\Gamma}) \ar@{->>}[rr] \ar@{->>}[rd] &
& A_{p}(k)
\\
& \U(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}) \ar@{->>}[ru] & }
\]
\end{proposition}
\noindent\textbf{Proof.}
This is a direct consequence of the surjective morphism of graded algebras given by $\YM(n,s)^{\Gamma} \rightarrow \YM(n)$ described at the end of
Subsection \ref{subsec:defi} and \cite{HS10}, Coro. 4.5.
\qed
On the other hand, we also have the:
\begin{theorem}
\label{teo:yangmillsclifweyl}
Let $n, s, p, q \in \NN$ be positive integers, satisfying that $n \geq 3$ and $s \geq 1$.
We suppose further that either $p \geq 3$, or $p = 2$ and $q \geq 2$.
Then, there exists a surjective homomorphism of super algebras
\[ \U(\ym(n,s)^{\Gamma}) \twoheadrightarrow \Cliff_{q}(k) \otimes A_{p}(k). \]
Furthermore, there exists $l \in \NN$ such that we can choose this morphism in such a way that it factors through the quotient
$\U(\ym(n,s)^{\Gamma}/F^{l}(\ym(n,s)^{\Gamma}))$
\[
\xymatrix@C-20pt@R-10pt { \U(\ym(n,s)^{\Gamma}) \ar@{->>}[rr] \ar@{->>}[rd] &
& \Cliff_{q}(k) \otimes A_{p}(k)
\\
& \U(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}) \ar@{->>}[ru] & }
\]
\end{theorem}
\noindent\textbf{Proof.}
We know that $\ym(n,s)^{\Gamma} = V(2)_{0} \oplus \hat{\tym}(n,s)^{\Gamma}$ as graded vector spaces.
As we have proved in Theorem \ref{teo:libre}, the Lie ideal $\hat{\tym}(n,s)^{\Gamma}$, considered as a graded Lie algebra, is a free graded Lie algebra
generated by a graded vector space $\hat{W}(n,s)^{\Gamma}$, that is,
\[ \hat{\tym}(n,s)^{\Gamma} \simeq \f(\hat{W}(n,s)^{\Gamma}). \]
We also point out that, by the computation of the Hilbert series of $\hat{W}(n,s)^{\Gamma}$, this space has an infinite number of
linearly independent even and odd elements (for $n \geq 3$ and $s \in \NN$).
We introduce the following notation, that we will need in the proof.
Given a super vector space $X$ with a nondegenerate (even) superantisymmetric bilinear form $\Omega$, consider the super vector space
$\mathfrak{heis}(X) = X \oplus k.z$.
It is a super Lie algebra provided with the bracket given by declaring that $z$ is central and $[x,x'] = \Omega(x,x') z$,
and it is called the \emph{Heisenberg super Lie algebra} defined by $(X,\Omega)$.
It is uniquely determined by the super dimension $(d,d')$ of $X$, so it will also be denoted by $\mathfrak{heis}_{d,d'}$.
More concretely, if the super dimension of $X$ is $(2r, 2t')$, $\mathfrak{heis}_{2r, 2t'}$ is the super Lie algebra with basis given by even elements
$q_{1}, \dots, q_{r}, p_{1}, \dots, p_{r}, z$ and odd ones $a_{1}, \dots, a_{t'}, b_{1}, \dots, b_{t'}$, such that $z$ is central,
$[q_{i},p_{j}] = \delta_{i,j} z$, $[a_{i},b_{j}] = \delta_{i,j} z$ and the other brackets vanish.
If the super dimension of $X$ is $(2r, 2t'+1)$, $\mathfrak{heis}_{2r, 2t'+1}$ is the super Lie algebra with basis given by even elements
$q_{1}, \dots, q_{r}, p_{1}, \dots, p_{r}, z$ and odd ones $a_{1}, \dots, a_{t'}, b_{1}, \dots, b_{t'},c$, such that $z$ is central,
$[q_{i},p_{j}] = \delta_{i,j} z$, $[a_{i},b_{j}] = \delta_{i,j} z$, $[c,c] = z$ and all other brackets vanish.
It is easy to see that $\cl{z-1} \subseteq \U(\mathfrak{heis}_{2r,t})$ is primitive, and in fact
$\U(\mathfrak{heis}_{2r,t})/\cl{z-1} \simeq \Cliff_{t}(k) \otimes A_{r}(k)$ (see \cite{BM89}, 0.2, (a) and (b)).
Denote the projection induced by the previous quotient by $\Pi_{\mathfrak{heis}_{2r,t}}$.
Hence, it suffices to prove the theorem for $\U(\mathfrak{heis}_{2p,q})$ instead of $\Cliff_{q}(k) \otimes A_{p}(k)$.
A morphism of super Lie algebras from $\hat{\tym}(n,s)^{\Gamma}$ to $\mathfrak{heis}_{2r,t}$ induces a respective morphism of super algebras
from $\U(\hat{\tym}(n,s)^{\Gamma})$ to $\U(\mathfrak{heis}_{2r,t})$.
Since $\hat{\tym}(n,s)^{\Gamma}$ is free with space of generators $\hat{W}(n,s)^{\Gamma}$, the former is equivalent to give a morphism of
super vector spaces from $\hat{W}(n,s)^{\Gamma}$ to $\mathfrak{heis}_{2r,t}$.
The morphism of super algebras will be surjective if the image of the corresponding morphism of super vector spaces is also so.
Since $\mathfrak{heis}_{2r,t}$ has finite super dimension, but $\hat{W}(n,s)^{\Gamma}$ has an infinite number of even and odd generators, this can be easily done.
From now on, we shall exclusively work in the super case (for algebras).
However, we will also keep track of the $\NN$-grading of the super Yang-Mills algebra $\ym(n,s)^{\Gamma}$, its ideal $\hat{\tym}(n,s)^{\Gamma}$
and the generator space $\hat{W}(n,s)^{\Gamma}$.
We shall now suppose that either $r \geq 1$, or $r = 0$ and $t \geq 2$, and show that under any of these assumptions, there exists a
$k$-linear homogeneous morphism (of super vector spaces) of degree $0$
\[ \phi : \hat{W}(n,s)^{\Gamma} \rightarrow \mathfrak{heis}_{2r,t} \]
such that there exists a set of linearly independent homogeneous elements of $\hat{W}(n,s)^{\Gamma}$
which are mapped onto a set of homogeneous generators of the Heisenberg super Lie algebra, respecting the degree.
$x_{3}$ is mapped to zero, $x_{13}$ and $x_{23}$ are mapped to a linearly independent set of two even generators of $\mathfrak{heis}_{2r,t}$,
and in fact
If $r \geq 1$, we set $\phi$ such that $\phi(x_{3}) = 0$, $\phi(x_{13}) = p_{1}$ and $\phi(x_{23}) = q_{1}$.
We also assume that for the set of other basis elements of even parity of the Heisenberg super Lie algebra $z$, $p_{i}$ and $q_{i}$, for $2 \leq i \leq r$,
there exist a linearly independent set of even homogeneous elements of degree greater than or equal to $6$,
$w_{i} \in \hat{W}(n,s)^{\Gamma}$ whose image under $\phi$ give the previous elements.
Let $d_{i}$ be the degree of $w_{i}$, for each ($2 \leq i \leq r$).
Let $j$ be the maximum between $4$ and the degrees $d_{i}$.
We moreover assume that there also exist linearly independent even (resp. odd) homogeneous elements of $\hat{W}(n,s)^{\Gamma}$ of degree greater than $j$
mapped onto a set of even (resp. odd) generators of the super Lie algebra $\mathfrak{heis}_{2r,t}$.
Let us called the maximum of these degrees by $d'$.
These conditions are easily verified, taking into account that $\hat{W}(n,s)^{\Gamma}$ has infinite dimensional even and odd components.
On the other hand, if $r = 0$ and $t \geq 2$, consider that $\phi(z_{1}) = 0$, $\phi([x_{1},z_{1}]) = a_{1}$ and $\phi([x_{2},z_{1}]) = b_{1}$,
where $a_{1}, b_{1} \in \mathfrak{heis}_{2r,t}$ are two linearly independent elements.
We also assume that for the set of other basis elements of the Heisenberg super Lie algebra there exist a linearly independent
set of even homogeneous elements of degree greater than or equal to $6$, $w_{i} \in \hat{W}(n,s)^{\Gamma}$,
such that each of the basis elements of the Heisenberg super Lie algebra is the image under $\phi$ of a respective element $w_{i}$ of the same degree.
Let $d_{i}$ be the degree of $w_{i}$, and let $j$ be the maximum between $4$ and the degrees $d_{i}$.
Again, these conditions are easily verified, taking into account that $\hat{W}(n,s)^{\Gamma}$ has infinite dimensional even and odd components.
Note that in either case there are a lot of choices for this morphism $\phi$.
Set $l$ to be $2 d' + 1$.
The map $\phi$ induces a unique surjective homomorphism $\Phi : \U(\hat{\tym}(n,s)^{\Gamma}) \twoheadrightarrow \U(\mathfrak{heis}_{2r,t})$,
equivalent to the homomorphism of super Lie algebras
\[ \hat{\tym}(n,s)^{\Gamma} \rightarrow \mathfrak{heis}_{2r,t}. \]
Since $\mathfrak{heis}_{2r,t}$ is nilpotent, the last morphism may be factorized in the following way
\[ \hat{\tym}(n,s)^{\Gamma} \rightarrow \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma} \rightarrow \mathfrak{heis}_{2r,t}, \]
where the first morphism is the canonical projection.
Hence, the map $\Phi$ may be factorized as
\[ \U(\hat{\tym}(n,s)^{\Gamma}) \twoheadrightarrow \U(\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}) \twoheadrightarrow \U(\mathfrak{heis}_{2r,t}). \]
We have thus obtained a surjective homomorphism of super algebras
\[ \Psi : \U(\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}) \twoheadrightarrow \U(\mathfrak{heis}_{2r,t}), \]
where the super Lie algebra $\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$ is obviously nilpotent and finite dimensional.
Moreover, it is a Lie ideal of the nilpotent super Lie algebra $\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$.
We have, as graded vector spaces,
\[ \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma} = V(2)_{0} \oplus \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}. \]
Let $I$ be the kernel of $\Pi_{\mathfrak{heis}_{2r,t}} \circ \Psi$ in $\U(\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$.
Taking into account that the quotient of $\U(\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$ by $I$ is a Clifford-Weyl super algebra which is simple, $I$ is a maximal two-sided ideal, and then,
there exists an even linear functional
$f \in (\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})^{*}$ such that $I = I(f)$.
We fix a polarization $\h_{f}$ of $\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$ at $f$.
Let $\bar{f} \in (\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})^{*}$ be any (even) extension of $f$.
Since $\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$ is an ideal of the nilpotent super Lie algebra
$\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$, by \cite{Her10}, Prop. 4.16, we have that the stabilizer
$\mathfrak{st}(I(f), \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$ includes
\[ \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma} + (\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})', \]
and we recall that $(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})'$ is given by
\[ \{ x \in \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma} : f([x , \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}]) = 0 \}. \]
If $r \geq 1$, we get immediately that $\bar{x}_{3} \in I$, but $\bar{x}_{13}$ and $\bar{x}_{23}$ do not belong to $I$,
since $\Psi(\bar{x}_{13})$ and $\Psi(\bar{x}_{23})$ do not vanish (they are in fact linearly independent).
Analogously, if $r = 0$ and $t \geq 2$, $\bar{z}_{1} \in I$, but $[\bar{x}_{1},\bar{z}_{1}]$ and $[\bar{x}_{2},\bar{z}_{1}]$ do not belong to $I$,
since $\Psi([\bar{x}_{1},\bar{z}_{1}])$ and $\Psi([\bar{x}_{2},\bar{z}_{1}])$ are linearly independent.
Let $x \in \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$, then $x = x' + y$, where
\[ x' = \sum_{i=1}^{2} c_{i} \bar{x}_{i} \in V(2)_{0}, \]
and $y \in \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$.
Since $[y,I(f)] \subset I(f)$, this implies that $x \in \mathfrak{st}(I(f), \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$ if and only if
$[x' , I(f)] \subset I(f)$.
In particular, if $r \geq 1$,
\[ [x' , \bar{x}_{3}] = \sum_{i=1}^{2} c_{i} [\bar{x}_{i} , \bar{x} _{3}]. \]
If $[x' , \bar{x}_{3}] \in I$, then $\Psi([x' , \bar{x}_{3}]) = 0$, or,
\[ \sum_{i=1}^{2} c_{i} \Psi(\bar{x}_{i3}) = 0, \]
but since $\Psi(\bar{x}_{13})$ and $\Psi(\bar{x}_{23})$ are linearly independent, we get that $c_{i} = 0$, for all $i = 1, 2$, which gives $x' = 0$.
In a similar way, if $r = 0$ and $t \geq 2$,
\[ [x' , \bar{z}_{1}] = \sum_{i=1}^{2} c_{i} [\bar{x}_{i} , \bar{z} _{1}], \]
so the assumption that $[x' , \bar{z}_{1}] \in I$, implies that
\[ \sum_{i=1}^{2} c_{i} \Psi([\bar{x}_{i} , \bar{z} _{1}]) = 0, \]
but since $\Psi([\bar{x}_{1} , \bar{z} _{1}])$ and $\Psi([\bar{x}_{2} , \bar{z} _{1}])$ are linearly independent, we get that $c_{i} = 0$, for all $i = 1, 2$, which again tells us that $x' = 0$.
In any case, we get that $\mathfrak{st}(I(f), \ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}) = \hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$.
So, the space $(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})'$ is included in the quotient $\hat{\tym}(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$,
and, by definition, $\h_{f}$ is also a polarization of $\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma}$ at $f$.
In particular, the weight of the ideal $I(\bar{f})$ can be computed easily using \cite{Her10}, Prop. 4.13, previously mentioned, which gives
$(r+2,t)$.
Therefore, the quotient of $\YM(n,s)^{\Gamma}$ by the inverse image of $I(\bar{f})$ under the projection
$\U(\ym(n,s)^{\Gamma}) \rightarrow \U(\ym(n,s)^{\Gamma}/F^{l}\ym(n,s)^{\Gamma})$ is isomorphic to $\Cliff_{t}(k) \otimes A_{r+2}(k)$ and the theorem follows.
\qed
\begin{remark}
We may also study the simplest case cases $\ym(1,1)^{\Gamma}$ and $\ym(1,2)^{\Gamma}$, which are nilpotent and finite dimensional,
as explained in Paragraph \ref{subsec:tym}.
Their representation theory can be thus directly analyzed using the Kirillov orbit method.
In particular, since $\ym(1,1)^{\Gamma}$ is a supercommutative, the only irreducible representations are one-dimensional.
Concerning the super Lie algebra $\ym(1,2)^{\Gamma}$, it is easy to check that for any even functional $f \in (\ym(1,2)^{\Gamma}_{0})^{*}$,
we have that $(\ym(1,2)^{\Gamma})^{f} = \ym(1,2)^{\Gamma}$ if $f = 0$, and $(\ym(1,2)^{\Gamma})^{f} = \ym(1,2)^{\Gamma}_{0}$, else.
This implies that the simple quotients of $\YM(1,2)^{\Gamma}$ are isomorphic to either $k$ or $\Cliff_{2}(k)$.
\end{remark}
\begin{remark}
\label{rem:otrosn}
Using a similar proof to the one given before, one can also show that, if $n = 1$ and $s \geq 3$, or $n = 2$, then there exists an infinite set of indices
$(p,q)$ such that
\[ \U(\ym(n,s)^{\Gamma}) \twoheadrightarrow \Cliff_{q}(k) \otimes A_{p}(k), \]
where $p \geq 2$ and $q \geq 4$, for $n = 1$ and $s \geq 3$; and $p \geq 4$ and $q \geq 3$, for $n \geq 2$.
Furthermore, there exists $l \in \NN$ such that we can select this morphism in such a way that it factors through the quotient
$\U(\ym(n,s)^{\Gamma}/F^{l}(\ym(n,s)^{\Gamma}))$.
The previous set can be chosen in such a way that if $(p,q)$ is not in it, then both $(p,q+1)$ and $(p,q-1)$ are.
\end{remark}
\begin{remark}
\label{rem:morita}
We would like to make some comments on the (abelian) category of representations of the Clifford-Weyl super algebras.
We remark that the category of representations of a super algebra $A$ is provided with homogeneous $A$-linear morphisms of degree zero.
Note that, even if a super algebra $A$ is concentrated in degree zero, its category of representations, as a super algebra, does not coincide with the category of representations of the underlying algebra of $A$.
Indeed, we remark that a representation of a super algebra $A$ concentrated in degree zero is given by a direct sum $M_{0} \oplus M_{1}$ of two modules $M_{0}$ and $M_{1}$ over the underlying algebra of $A$,
and a morphism from $M_{0} \oplus M_{1}$ to $N_{0} \oplus N_{1}$ is given by a pair $(f_{0},f_{1})$, where $f_{i} : M_{i} \rightarrow N_{i}$ is a morphism of modules over the underlying algebra of $A$,
for all $i \in \ZZ/2\ZZ$.
Two super algebras $A$ and $B$ are called \emph{Morita equivalent} if there is an equivalence between the categories of representations of the super algebra $A$ and the super algebra $B$, which commutes with the shift.
We remark that if two super algebras are Morita equivalent, then the underlying algebras are also Morita equivalent
(\textit{cf.} \cite{GG82}, Section 5, where the authors work with graded algebras, but the statements are analogous, in particular: Lemma 5.1, Cor. 5.2, Prop. 5.3 and Thm. 5.4).
We now note that the super algebra $\Cliff_{2q}(k) \simeq M_{2^{q}}(k)$ is Morita equivalent to the super algebra $k$.
This implies that the Clifford-Weyl super algebra $\Cliff_{2q}(k) \otimes A_{p}(k)$ is Morita equivalent to the Weyl algebra $A_{p}(k)$,
which is regarded as a super algebra concentrated in degree zero, and the super algebras $\Cliff_{2q + 1}(k) \otimes A_{p}(k)$ and
$\Cliff_{1}(k) \otimes A_{p}(k)$ are also Morita equivalent.
Moreover, a direct inspection tells us that the category of representations of the super algebra $\Cliff_{1}(k) \otimes A_{p}(k)$ is equivalent to the category of representations of the algebra $A_{p}(k)$.
We see thus that in either case the representations that we obtain can be understood as induced by those of the Weyl algebras.
We would also like to point out that several families of representations of the Weyl algebras have been previously studied by Bavula and Bekkert in \cite{BB00}, and, by the theorem,
they can also be also used to induce representations of the super Yang-Mills algebras.
\end{remark}
\begin{bibdiv}
\begin{biblist}
\bib{AL85}{article}{
author={Aubry, Marc},
author={Lemaire, Jean-Michel},
title={Zero divisors in enveloping algebras of graded Lie algebras},
journal={J. Pure Appl. Algebra},
volume={38},
date={1985},
number={2-3},
pages={159--166},
}
\bib{AZ94}{article}{
author={Artin, M.},
author={Zhang, J. J.},
title={Noncommutative projective schemes},
journal={Adv. Math.},
volume={109},
date={1994},
number={2},
pages={228--287},
}
\bib{BB00}{article}{
author={Bavula, Vladimir},
author={Bekkert, Viktor},
title={Indecomposable representations of generalized Weyl algebras},
journal={Comm. Algebra},
volume={28},
date={2000},
number={11},
pages={5067--5100},
}
\bib{BM89}{article}{
author={Bell, Allen D.},
author={Musson, Ian M.},
title={Primitive factors of enveloping algebras of nilpotent Lie
superalgebras},
journal={J. London Math. Soc. (2)},
volume={42},
date={1990},
number={3},
pages={401--408},
}
\bib{Ber08}{article}{
title={La cat\'egorie des modules gradu\'es sur une alg\`ebre gradu\'ee},
author={Berger, Roland},
journal={Preprint},
date={2008},
pages={1--21}
}
\bib{BG06}{article}{
title={Higher symplectic reflection algebras and non-homogeneous $N$-Koszul property},
author={Berger, Roland},
author={Ginzburg, Victor},
journal={J. Algebra},
volume={304},
number={1},
date={2006},
pages={577--601}
}
\bib{BM06}{article}{
author={Berger, Roland},
author={Marconnet, Nicolas},
title={Koszul and Gorenstein properties for homogeneous algebras},
journal={Algebr. Represent. Theory},
volume={9},
date={2006},
number={1},
pages={67--97},
}
\bib{BT07}{article}{
author={Berger, Roland},
author={Taillefer, Rachel},
title={Poincar\'e-Birkhoff-Witt deformations of Calabi-Yau algebras},
journal={J. Noncommut. Geom.},
volume={1},
date={2007},
number={2},
pages={241--270},
}
\bib{Berg78}{article}{
author={Bergman, George M.},
title={The diamond lemma for ring theory},
journal={Adv. in Math.},
volume={29},
date={1978},
number={2},
pages={178--218},
}
\bib{CM84}{article}{
author={Cohen, M.},
author={Montgomery, S.},
title={Group-graded rings, smash products, and group actions},
journal={Trans. Amer. Math. Soc.},
volume={282},
date={1984},
number={1},
pages={237--258},
}
\bib{CD02}{article}{
author={Connes, Alain},
author={Dubois-Violette, Michel},
title={Yang-Mills algebra},
journal={Lett. Math. Phys.},
volume={61},
date={2002},
number={2},
pages={149--158},
}
\bib{CD07}{article}{
author={Connes, Alain},
author={Dubois-Violette, Michel},
title={Yang-Mills and some related algebras},
conference={
title={Rigorous quantum field theory},
},
book={
series={Progr. Math.},
volume={251},
publisher={Birkh\"auser},
place={Basel},
},
date={2007},
pages={65--78},
}
\bib{DF99}{article}{
author={Deligne, Pierre},
author={Freed, Daniel S.},
title={Classical field theory},
conference={
title={},
address={Princeton, NJ},
date={1996/1997},
},
book={
publisher={Amer. Math. Soc.},
place={Providence, RI},
},
date={1999},
pages={137--225},
}
\bib{DF99s}{article}{
author={Deligne, Pierre},
author={Freed, Daniel S.},
title={Supersolutions},
conference={
title={},
address={Princeton, NJ},
date={1996/1997},
},
book={
publisher={Amer. Math. Soc.},
place={Providence, RI},
},
date={1999},
pages={227--355},
}
\bib{FKM83}{article}{
author={Ferrero, Miguel},
author={Kishimoto, Kazuo},
author={Motose, Kaoru},
title={On radicals of skew polynomial rings of derivation type},
journal={J. London Math. Soc. (2)},
volume={28},
date={1983},
number={1},
pages={8--16},
}
\bib{Fr75}{article}{
author={Fr{\"o}berg, Ralph},
title={Determination of a class of Poincar\'e series},
journal={Math. Scand.},
volume={37},
date={1975},
number={1},
pages={29--39},
}
\bib{GG82}{article}{
author={Gordon, Robert},
author={Green, Edward L.},
title={Graded Artin algebras},
journal={J. Algebra},
volume={76},
date={1982},
number={1},
pages={111--137},
}
\bib{HKL08}{article}{
author={Hai, Ph{\`u}ng H{\^o}},
author={Kriegk, Benoit},
author={Lorenz, Martin},
title={$N$-homogeneous superalgebras},
journal={J. Noncommut. Geom.},
volume={2},
date={2008},
number={1},
pages={1--51},
}
\bib{Her10}{article}{
author={Herscovich, Estanislao},
title={The Dixmier map for nilpotent super Lie algebras},
journal={Preprint},
date={2010},
eprint={http://arxiv.org/pdf/1009.1124},
}
\bib{HS10}{article}{
author={Herscovich, Estanislao},
author={Solotar, Andrea},
title={Representation theory of Yang-Mills algebras},
journal={Ann. of Math. (2)},
volume={173},
number={2},
publisher={Springer-Verlag},
date={2011},
pages={1043--1080},
}
\bib{HS10h}{article}{
author={Herscovich, Estanislao},
author={Solotar, Andrea},
title={Hochschild and cyclic homology of Yang-Mills algebras},
journal={Accepted for publication in Journal f\"ur die Reine und Angewandte Mathematik},
date={2010},
}
\bib{Lang02}{book}{
author={Lang, Serge},
title={Algebra},
series={Graduate Texts in Mathematics},
volume={211},
edition={3},
publisher={Springer-Verlag},
place={New York},
date={2002},
}
\bib{Let89}{article}{
author={Letzter, Edward},
title={Primitive ideals in finite extensions of Noetherian rings},
journal={J. London Math. Soc. (2)},
volume={39},
date={1989},
number={3},
pages={427--435},
}
\bib{MM11}{article}{
author={Mori, Izuru},
author={Minamoto, Hiroyuki},
title={The structure of AS-Gorenstein algebras},
journal={Adv. Math.},
volume={226},
date={2011},
number={5},
pages={4061--4095},
}
\bib{Mov05}{article}{
author={Movshev, Michael},
title={Yang-Mills theories in dimensions 3,4,6,10 and Bar-duality},
journal={Preprint},
eprint={http://arxiv.org/abs/hep-th/0503165v2},
date={2005},
pages={1--47},
}
\bib{MS06}{article}{
author={Movshev, Michael},
author={Schwarz, Albert},
title={Algebraic structure of Yang-Mills theory},
conference={
title={The unity of mathematics},
},
book={
series={The unity of Mathematics, Progr. Math.},
volume={244},
publisher={Birkh\"auser Boston},
place={Boston, MA},
},
date={2006},
pages={473--523},
}
\bib{MPU09}{article}{
author={Musson, Ian M.},
author={Pinczon, Georges},
author={Ushirobira, Rosane},
title={Hochschild cohomology and deformations of Clifford-Weyl algebras},
journal={SIGMA Symmetry Integrability Geom. Methods Appl.},
volume={5},
date={2009},
pages={Paper 028, 27},
}
\bib{Pi08}{article}{
author={Piontkovski, Dmitri},
title={Coherent algebras and noncommutative projective lines},
journal={J. Algebra},
volume={319},
date={2008},
number={8},
pages={3280--3290},
}
\bib{Rot09}{book}{
author={Rotman, Joseph J.},
title={An introduction to homological algebra},
series={Universitext},
edition={2},
publisher={Springer},
place={New York},
date={2009},
}
\bib{Ta95}{article}{
author={Tanaka, Junko},
title={On homology and cohomology of Lie superalgebras with coefficients
in their finite-dimensional representations},
journal={Proc. Japan Acad. Ser. A Math. Sci.},
volume={71},
date={1995},
number={3},
pages={51--53},
}
\bib{Wei94}{book}{
title={An introduction to homological algebra},
author={Weibel, Charles A.},
publisher={Cambridge University Press},
place={Cambridge},
series={Cambridge Studies in Advanced Mathematics},
volume={38},
date={1994},
}
\end{biblist}
\end{bibdiv}
\noindent Estanislao Herscovich,
\\
Fakult\"at f\"ur Mathematik,
\\
Universit\"at Bielefeld,
\\
D-33615 Bielefeld,
\\
Germany,
\\
\href{mailto:<EMAIL>}{<EMAIL>}
\end{document}
|
\section{Introduction}
\indent
Although morphemes, not letters, are usually considered to be
the smallest linguistic unit, studying statistics of letter usage
has its own merit. For example, information on letter frequency is
essential in cryptography for deciphering a substitution code \citep{friedman},
and ``frequency analysis" was used in as early as the 9th century
by the Arab scientist al-Kindi for the purpose of decryption \citep{al-kindi}.
An efficient design of a communication code also depends
crucially on the letter frequency. The shortest Morse
code is reserved to letters that are the most common:
one dot for letter {\sl e} and one dash for letter {\sl t},
both letters being the most frequent in English.
The same principle is also behind the design of minimum-redundancy
code by Huffman \citep{huffman}.
The initial motivation for the ``QWERTY" mechanical typewriter design is to keep the most
common letters far away in the keyboard so that metal bars would not
jam for a fast typist \citep{david}. Even in modern times,
the digraph (letter pairs) frequency is an important piece
of information for keyboard design \citep{zhai}.
In all these examples, a quantitative description of
letter usage frequency is important. Unlike the ranked word
frequency distribution, which is well characterized by a simple
power-law function or Zipf's law \citep{zipf}, it is not
clear whether a universal fitting function exists despite
a claim of such a function (the logarithmic function) in \citep{kanter}.
In this paper, we aim at critically examining various
functional forms of fitting rank-frequency distribution
of letters, ranging from simple to more complicated
ones with two or three free parameters. The dataset used is
the historical U.S. and Mexican presidential speeches.
The presidential speeches are readily available
(see another study where the Italian presidential ``end of year"
addresses are used \citep{tuzzy}), they also offer an
opportunity for investigations of temporal patterns in
letter usage.
The ranked word frequency distributions studied by George Zipf
have extremely long tails, due to the presence of
low-frequency words (such as {\sl hapax legomena}). As a result,
logarithmic transformation is usually applied to the $x$-axis
(as well as the $y$-axis). The double logarithmic transformation
is also justified by the expectation of a power-law
function, as it will lead to a linear regression.
This linear fitting in log-log scale may have its pitfall
\citep{newman}, one being the uneven distribution of points
along the log-transformed $x$-axis.
For ranked letter frequency data, the finite number of
alphabets sets an upper bound for the rank, and there
is no large number of rare events which is an important
theoretical issue in modeling the word rank-frequency
distribution \citep{baayen}. On the other hand, the
limited range of abscissa may make it hard to distinguish
different fitting functions. Since power-law function
is not expected to be the best fitting function,
double logarithmic transformation is not necessary,
and we will fit the data in linear-linear scale.
No longer linear fittings, the curve fitting is
carried out by nonlinear least-square \citep{nls}.
Statistical models with a larger number of free parameters will
guarantee to fit the data better than a model subset with fewer
number of parameters. To compare the performance of models
with different number of parameters, a penalty should be
imposed on the extra number of parameters. Towards this end,
we apply the standard model selection technique with
Akaike Information criterion or AIC \citep{aic,aic-book} and
Bayesian Information Criterion or BIC \citep{bic} to compare
various functions used to fit the ranked letter frequency data.
\section{Data}
\indent
{\bf US presidential inaugural speeches:}
In order to take into account of any possible letter usage
trend in time, we use the US Presidential Inaugural
Speech texts for the 44 presidents in the last 200 years.
The data is downloaded from the {\sl The American Presidency
Project} from the University of California at Santa Barbara
site ({\sl http://www.presidency.ucsb.edu/}).
Multiple inaugural speeches from the same person are combined
into one, including the nonconsecutive presidency of
Grover Cleveland. Five presidents did not give an inaugural
speech (John Tyler, Millard Fillmore, Andrew Johnson, Chester Arthur, Gerald Ford).
The final dataset consists of 38 text files.
{\bf Mexico presidential addresses to the congress:}
For Spanish texts, we selected the 19 Mexican presidents'
report to congress ({\sl Informes Presidenciales}) from 1914 to 2006.
\\
({\sl http://www.diputados.gob.mx/cedia/sia/re\_info.htm} )
Again, addresses by the same president are combined into
one text file. Some presidential texts are much shorter than others
due to two possible reasons: either did the president
only present one address (the typical number of addresses is 6),
such as Adolfo de la Huerta (1920) and Emilio Portes Gil (1929),
or the president gave shorter reports, such as Ernesto
Zedillo Ponce de Le\'{o}n (1995-2000) and Vicente Fox Quesada (2001-2006).
\section{Letter frequencies and their temporal trends}
\indent
Fig.\ref{fig1}(A) shows the English letter frequency of the 38 US president's speeches,
separated by the century. The letter {\sl e} remains the most
commonly used English letter with little change in its frequency.
However there seems to be a trend of less usage of letter {\sl t},
and more usage of letter {\sl w} in the 20th century as compared to
the 19th century.
\begin{figure}[th]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig1-letter-freq-by-period.eps, width=8.5cm}
\end{turn}
\end{center}
\caption{
\label{fig1}
English (A) and Spanish (B) letter frequencies (unranked,
in alphabetic order) for 38 U.S. presidential inaugural
speeches and 19 Mexican presidents' report to congress.
Letter frequencies of each president's speech are linked
by a line, and different time periods are drawn separately
(U.S. president speech: 1789-1800, shifting 1801-1901 by
0.02, shifting 1905-now by 0.04; Mexico president speech:
1919-1934, shifting 1935-1964 by 0.02, shifting 1965-2005
by 0.04). Due to a larger sample size, the fluctuation of
frequency from president to president in Spanish texts is much smaller
than that in English texts.
}
\end{figure}
Similar frequencies of Spanish letters in the 19 Mexican presidents'
addresses are shown in Fig.\ref{fig1}(B). Letters with accent
(the acute accent for the vocals, the umlaut for
the letter \"{u} and the tilde for the \~{n})
are counted separatedly but later combined because they do not
really represent different letters.
The 19 files are arbitrarily split into three groups:
the first 7 presidents (from 1917 to 1934), the next 5 presidents
(from 1935 to 1964), and the last 7 presidents (from 1965 to 2006).
These three groups are separatedly drawn in Fig.\ref{fig1}(B).
The narrowing of the variations of letter frequency in Fig.\ref{fig1}(B)
as compared to Fig.\ref{fig1}(A) is due to the larger sample sizes
in Spanish texts.
\begin{table}
\small
\begin{tabular}{ccccc}
\hline
& name & year(s) of speech & sorting of alphabets & num \\
\hline
1 & Washington & 1789,1793 & etinoashrcdlumfpybwgvxjqkz & 7710 \\
2 & Adams & 1797 & etnioasrhdlcfumpgybvwxjkqz & 11281 \\
3 & Jefferson & 1801,1805 & etoinasrhldcufmpwgybvkxjzq & 18701 \\
4 & Madison & 1809,1813 & etoinasrhldcufpmgwbyvkxjzq & 11572 \\
5 & Monroe & 1817,1821 & etinoarshdclufpmwygbvxkjzq & 37522 \\
6 & Adams & 1825 & etoinasrhdlcfupmgybvwxjqkz & 14572 \\
7 & Jackson & 1829,1833 & etoinarshldcufmpybgwvxkjqz & 11372 \\
8 & Van Buren & 1837 & etoinasrhlducfpmywgvbxkjqz & 19215 \\
9 & Harrison & 1841 & etoinarshcdlfumpygbwvxkjzq & 40526 \\
10 & Polk & 1845 & etoniasrhdlcufmpygbwvxjqkz & 23475 \\
11 & Taylor & 1849 & etoinasrhlcdufmpybgwvxkjzq & 5413 \\
12 & Pierce & 1853 & etinoarshlducfmpygwbvkxjqz & 16406 \\
13 & Buchanan & 1857 & etionasrhlcdufpmywgvbxqjkz & 13696 \\
14 & Lincoln & 1861,1865 & etoinasrhldcufpmywbgvkxjqz & 19340 \\
15 & Grant & 1869,1873 & etoinarshldcufmpgywbvxkqjz & 11476 \\
16 & Hayes & 1877 & etoinasrhlcdufpmygbwvjqxkz & 12171 \\
17 & Garfield & 1881 & etoniasrhlducfmpgwybvkxjqz & 14477 \\
18 & Cleveland & 1885 & etoinasrhdlcufpmygbwvxkzjq & 18480 \\
19 & Harrison & 1889 & etoniasrhldcufpmwygbvkxjqz & 21394 \\
20 & Mckinley & 1897,1901 & etnoiarshlducfpmygbwvxkjzq & 30179 \\
21 & T.Roosevelt & 1905 & etoainrshldufwcgbpmvykxjzq & 4480 \\
22 & Taft & 1909 & etoinasrhdclfumpgywbvkxjqz & 26272 \\
23 & Wilson & 1913,1917 & etoanisrhdlucfwpmgyvbkjqxz & 14360 \\
24 & Harding & 1921 & etnioarsldhcufmwpgybvkxzjq & 16508 \\
25 & Coolidge & 1925 & etonairshldcufmpwybgvxkjqz & 19482 \\
26 & Hoover & 1929 & etoinarshldcufmpgywbvzxjkq & 19256 \\
27 & F.D.Roosevelt
& 1933,1937,1941,1945 & etoainrshldcfumpwygvbkjxzq & 25696 \\
28 & Truman & 1949 & etoainrshldcfumpwgyvbkjqxz & 11070 \\
29 & Eisenhower & 1953,1957 & etoainrshldfcumpwygbvkqjxz & 18313 \\
30 & Kennedy & 1961 & etoanrsihldfuwcmgypbvkjxzq & 6003 \\
31 & Johnson & 1965 & etanoirshdluwcfmgybpvkjxzq & 6468 \\
32 & Nixon & 1969,1973 & etoanirshldcuwfmpgbyvkjqxz & 17142 \\
33 & Carter & 1977 & etaonirshldumwcfpgbyvkjqxz & 5459 \\
34 & Reagan & 1981, 1985 & etonarishdlumwcfgpybvkjxzq & 22494 \\
35 & G.H.W.Bush & 1989 & etaonrishdluwcmgfybpvkzjxq & 9781 \\
36 & Clinton & 1993,1997 & eotanrishldcumwfpgybvkjzxq & 16915 \\
37 & G.W.Bush & 2001,2005 & etonairsdhlcufmwygpbvkjzqx & 16759 \\
38 & Obama & 2009 & etoanrsihdlucwfmgypbvkjqxz & 10632 \\
\hline
\end{tabular}
\caption{
\label{table1}
The names of the 38 U.S. presidents, the years of their inaugural
speech, the order of letters ranked by their frequency in
the corresponding president's speech, and the total counts of letters.
}
\end{table}
In Table \ref{table1} English letters are sorted by their frequency of usage, from common
to rare, for the 38 US president speeches. Again, {\sl e} and {\sl t} are
consistently ranked as number 1 and 2 (with the exception of
Clinton's speech, where {\sl o} is ranked second), but the ranking
order of {\sl a } and {\sl i} seem to change with time:
in older speeches (e.g. before year of 1890), {\sl i} is
ranked higher than {\sl a}, after 10 more presidents where
{\sl i} and {\sl a} were used about equally, then the order is
reversed for newer speeches (e.g. after the year 1960).
\begin{table}[ht]
\begin{tabular}{ccccc}
\hline
& name & years of speech & sorting of alphabets & num \\
\hline
1 & Carranza & 1917,1918,1919 & eaosnirdlctupmbgyvfqhjxz\~{n}kw & 539107 \\
2 & De la Huerta & 1920 & eaosinrdlctupmbgfvyhqjzx\~{n}kw & 113057 \\
3 & Obreg\'{o}n & 1921,1922,1923,1924 & eaosinrdlctupmbgyfvhqjxz\~{n}kw & 675552 \\
4 & El\'{i}as & 1925,1926,1927,1928 & eaosinrdlctupmbgyfvqhjzx\~{n}kw & 700715 \\
5 & Portes Gil & 1929 & eaosinrdlctupmbgyvfqhjzx\~{n}kw & 231873 \\
6 & Ortiz & 1930,1931,1932 & eaoisnrdlctupmbgvfyqhjzx\~{n}kw & 664319 \\
7 & Rodr\'{i}uez & 1933,1934 & eaoisnrdlctupmbgyvfqhjzx\~{n}kw & 301745 \\
8 & C\'{a}rdenas & 1935,1936,1937,1938,1939,1940& eaosinrldctupmbgvyfqhjxz\~{n}kw & 402748 \\
9 & \'{A}vila & 1941,1942,1943,1944,1945,1946& eaosinrlcdtumpbygvfqhjzx\~{n}kw & 734540 \\
10 & Alem\'{a}n & 1947,1948,1949,1950,1951,1952& eaoisnrcltdumpbyvgfhqzjx\~{n}kw & 549980 \\
11 & Ruiz & 1953,1954,1955,1956,1957,1958& eaosinrldctumpbygvfqhjzx\~{n}kw & 592550 \\
12 & L\'{o}pez & 1959,1960,1961,1962,1963,1964& eaosinrldctupmbgvyfhqzjx\~{n}kw & 712056 \\
13 & D\'{i}az & 1965,1966,1967,1968,1969,1970& eaosinrldctupmbgvyfqhzjx\~{n}kw & 785528 \\
14 & Echeverr\'{i}a& 1971,1972,1973,1974,1975,1976& eaosinrldctumpbvgfyqhjzx\~{n}kw & 792338 \\
15 & L\'{o}pez Portillo
& 1977,1978,1979,1980,1981,1982& eaosinrlcdtumpbygvfqhzjx\~{n}kw & 684658 \\
16 & De la Madrid & 1983,1984,1985,1986,1987,1988& eaoisnrlcdtumpbyvgfhqzjx\~{n}kw & 761274 \\
17 & Salinas & 1989,1990,1991,1992,1993,1994& eaosinrlcdtumpbvygfhqzxj\~{n}kw & 624933 \\
18 & Zedillo & 1995,1996,1997,1998,1999,2000& eaosinrlcdtumpbgyvfqhzjx\~{n}kw & 282463 \\
19 & Fox & 2001,2002,2003,2004,2005 & eaosinrldctumpbgyvfqhzjx\~{n}kw & 311429 \\
\hline \\
\end{tabular}
\caption{
\label{table2}
The last names of the 19 Mexican presidents, the years when they addressed
the congress, the order of letters ranked by their frequency in
the corresponding president's address, and the total counts of letters.
}
\end{table}
\large
Table \ref{table2} shows the corresponding sorting of Spanish letters
in the 19 Mexico president addresses. The sequence
{\sl eaosinr} consistently appears at the head of
the string. However, the order of {\sl d} and {\sl l}
has been switched from {\sl dl} in the first half of 20th century
(until president Rodr\'{i}uez whose term ended in 1934)
to {\sl ld} in the second half of the century (since
president Alem\'{a}n whose term started in 1946).
\begin{figure}[th]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig2-trend-letter-color.eps, width=10cm}
\end{turn}
\end{center}
\caption{
\label{fig2}
Temporal change of frequency in selected letters.
(A) Letters {\sl t, i, a, w} in 38 U.S. presidential speeches.
(B) Letters {\sl d,l,m} in 19 Mexican presidential speeches.
}
\end{figure}
To confirm the observation from Fig.\ref{fig1} and Tables \ref{table1},\ref{table2},
in Fig.\ref{fig2} we directly plot the English letter frequencies of {\sl t,w,a,i}
and Spanish letter frequencies of {\sl d,l,m}.
Indeed, there is higher usage of {\sl w} and lesser usage
of {\sl t} in recent US president speeches, and the relative order
of {\sl a} and {\sl t} was switching from year 1889 to 1957.
For Mexican president addresses, the letter {\sl l}
overcomes {\sl d} in the last few decades. There is
also an upward trend for the usage of Spanish letter {\sl m}.
Despite these interesting trends of a few letters for the last
two hundreds of years for English and one hundred of
years in Spanish, the overall letter frequencies remain
more or less stable. We combine all 38 English files into one
(and 19 Spanish files into one) to examine the rank frequency
distribution.
\large
\section{Fitting ranked letter frequency distributions}
\indent
We used ten different functions to fit the ranked letter
frequency distribution in US presidential inaugural speeches that
is averaged over all 38 presidents, and Mexican presidential
addresses to the congress averaged over 19 presidents. Here is a list of these
functions ($f$ denotes the normalized letter frequency, $r$ denotes
the rank: $r=1$ for most frequent letter and $r=26$ (or 27)
for the rarest letter, and $n=26,27$ is the maximum rank value):
\begin{eqnarray}
\label{eq-all}
\mbox{Gusein-Zade} &:& f= C \log \frac{n+1}{r} \\
\mbox{power-law} &:& f= \frac{C}{r^a} \\
\mbox{exponential} &:& f= C e^{-ar} \\
\mbox{logarithmic} &:& f= C - a \log(r) \\
\mbox{Weibull} &:& f= C \left( \log \frac{n+1}{r}\right)^a \\
\mbox{quadratic logarithmic} &:& f= C - a \log(r) - b \left( \log(r) \right)^2 \\
\mbox{Yule} &:& f= C \frac{b^r}{r^a} \\
\mbox{Menzerath-Altmann/Inverse-Gamma} &:& f= C \frac{e^{-b/r}}{r^a} \\
\mbox{Cocho/Beta} &:& f= C \frac{ (n+1-r)^b}{r^a} \\
\mbox{Frappat} &:& f= C + br + c e^{-ar}
\end{eqnarray}
Since $f$ is the normalized frequency, $\sum_{i=1}^n f_i =1$, which
adds a constrain on one parameter. The parameter under constraint
is labeled as $C$ whose value is generally of no interest to us. Besides $C$,
the number of free (adjustable) parameters in these fitting functions
ranges from 0 (Gusein-Zade) to 3 (Frappat). The power-law, exponential,
logarithmic, and Weibull functions have 1 free parameter,
quadratic-logarithmic, Yule, Menzerath-Altmann/Inverse-Gamma, and Cocho/Beta,
functions have 2 free parameters, as discussed in \citep{entropy10}
The power-law (Eq.(2)) and exponential function (Eq.(3)) are
often the first group of function to be tested, due to their simplicity
and widespread applicability.
The zero-free-parameter function (Gusein-Zade) in Eq.(1) \citep{gusein1,gusein2,gusein3}
actually
corresponds to the exponential cumulative distribution,
and the Weibull function (Eq.(5)) \citep{nabeshima} corresponds to the
stretched exponential cumulative distribution. The conversion
from cumulative distribution to rank distribution of these two
functions are discussed in details in \citep{entropy10}.
The logarithmic function (Eq.(4)) is an extension of the Gusein-Zade
function $C\log(n+1)-C\log(r)$ by allowing the coefficient of
$\log(r)$ term to be independently fitted. Then the quadratic
logarithmic function is an extension of the logarithmic function
by adding one extra term. The logarithmic function is mentioned
in \citep{kanter,vlad}, whereas quadratic logarithmic function has
not been used to the best our knowledge.
The three two-parameter functions used are all attempts to
modify the power-law function: Yule function \citep{yule25,martindale} uses an
exponential function ($b^r$), Menzerath-Altmann or inverse-Gamma
function \citep{altmann} uses an exponential function
of the inverse of rank ($e^{-b/r}$), and Cocho or Beta function
\citep{beta1,beta2,beta3}
uses a power-law function of the reverse rank ($(n+1-r)^b$).
The 3-parameter function in Eq.(8) proposed in \citep{frap,frap2} is to add a linear
trend over the exponential function.
All $x$ and $y$ relationship in Eqs.(1-10) are non-linear. It is possible
to transform variables or introduce new variables to carry
out the fitting by multiple linear regression. For example,
after define $y'=\log(f)$, $x'_1=\log(r)$, $x'_2=\log(n+1-r)$,
the Cocho/Beta function is equivalent to a multiple
regression $y'= c_0 +c_1 x_1 +c_2 x_2$, where
the regression coefficients can be converted back to the
parameters used in Eq.(7): $C=e^{c_1}, a=-c_1, b=c_2 $.
The data-fitting result in the transformed variable, however,
is generally not identical to the result in its original
nonlinear form. Our method is to first use the multiple linear
regression in the transformed version, if possible, in
order to obtain a rough estimation of the parameter values.
Then these values are used as the initial condition for
nonlinear least-square iteration
(using the {\sl nls} function \citep{nls} in R:
{\sl http://www.r-project.org/}).
\begin{figure}[th!]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig3-ranked-english-letter-freq.eps, width=11cm}
\end{turn}
\end{center}
\caption{
\label{fig3}
Fitting ranked English letter frequency of U.S. presidential
speech by ten different functions:
(A) power-law ($a=0.616$) and exponential function ($a=0.118$);
(B) Gusein function ($C=0.0374$);
(C) logarithmic function ($a=0.0401$);
(D) Weibull function ($a=0.935$);
(E) quadratic logarithmic function ($a=0.0280$, $b=0.00325$);
(F) Yule function ($a=0.0543$, $b=0.897$);
(G) Menzerath-Altmann/Inverse-Gamma function ($a=-1.05$, $b=-1.31$);
(H) Cocho/Beta function ($a=0.210$, $b=1.35$);
(I) Frappat function ($a=0.245$, $b=-0.00242$, $c=0.0813$).
The fitting performance measured by SSE and AIC/BIC is shown
in Table \ref{table3}.
}
\end{figure}
\begin{figure}[th!]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig4-fitting-ranked-spanish-letter-freq.eps, width=11cm}
\end{turn}
\end{center}
\caption{
\label{fig4}
Fitting ranked Spanish letter frequency of Mexican presidents'
speech to congress by ten different functions:
(A) power-law ($a=0.653$) and exponential function ($a=0.130$);
(B) Gusein function ($C=0.0303$);
(C) logarithmic function ($a=0.0443$);
(D) Weibull function ($a=1.05$);
(E) quadratic logarithmic function ($a=0.0306$, $b=0.00362$);
(F) Yule function ($a=-0.0333$, $b=0.873$);
(G) Menzerath-Altmann/Inverse-Gamma function ($a=-1.22$, $b=-1.69$);
(H) Cocho/Beta function ($a=0.115$, $b=2.04$);
(I) Frappat function ($a=0.0592$, $b=0.00315$, $c=0.276$).
The fitting performance measured by SSE and AIC/BIC is shown
in Table \ref{table3}.
}
\end{figure}
\begin{figure}[th!]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig5-9-error-US.eps, width=9cm}
\end{turn}
\end{center}
\caption{
\label{fig5}
Fitting errors (residual, deviance), $y_{(r)}-f(r)$, of the
ten functions used in Fig.\ref{fig3} for U.S. presidential
speeches.
}
\end{figure}
\begin{figure}[th!]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig6-9-error-mexico.eps, width=9cm}
\end{turn}
\end{center}
\caption{
\label{fig6}
Fitting errors (residual, deviance), $y_{(r)}-f(r)$, of the
ten functions used in Fig.\ref{fig4} for Mexican presidents'
speech to congress.
}
\end{figure}
Fig.\ref{fig3} shows the nonlinear least-square fitting
of English letter ranked frequencies with all ten functions
in Eqs.(1-10), and Fig.\ref{fig4} shows the result for
Spanish ranked letter frequencies. The first impression
of Figs.\ref{fig3},\ref{fig4} is that all functions seem to fit
the ranked letter frequency well, with the exception of
power-law and Menzerath-Altmann functions. Is it possible to
further distinguish those with even better fitting
performance? That is the issue to be addressed in
the next section.
\section{Comparison of the fitting performance}
\indent
How well a function $f$ fits the data can be measured by
the sum of squared errors (residuals) SSE:
\begin{equation}
SSE= \sum_{i=1}^n (y_i - \hat{f}(x_i))^2
\end{equation}
where the parameters of the function are estimated by
least-square or maximum likelihood method. It is not
correct to compare two functions with different number of
parameters, as the function with more parameters has
more freedom to adjusting in order to achieve a higher
fitting performance. In the extreme example, a function with
unlimited number of parameters can fit a finite dataset
perfectly: this overfitting situation is called saturation.
To compare two functions with different number of
parameters, the Akaike Information Criterion (AIC)
\citep{aic} and Bayesian Information Criterion (BIC)
\citep{bic} can be used for model selection. Both criteria
discount the (log) maximum likelihood of the fitting
model by a term proportional to the number of parameters ($p$):
AIC uses the term 2$p$, and BIC uses the term $\log(n)p$
(where $n$ is the sample size). Maximizing the discounted
maximum likelihood is our criterion for the best model
(equivalent to minimizing AIC or BIC) \citep{aic-book}.
In regression models (linear or nonlinear), there is
a simple relationship between AIC/BIC and SSE if
we assume the variance of errors is unknown (and has
to be estimated from the data), and if we assume the
variance of the error is the same for all data points (details are
in Appendix).
\begin{table}[ht]
\begin{center}
\begin{tabular}{c|c|c|ccc|ccc}
\hline
function & Eq. & $p$ & \multicolumn{3}{c|}{English} &
\multicolumn{3}{c}{Spanish} \\
\cline{4-9}
& & & SSE & $\Delta$ AIC & $\Delta$ BIC & SSE & $\Delta$ AIC & $\Delta$ BIC \\
\hline
Gusein-Zade & 1 & 0 & 0.00106 & 20.2 & 17.7 & 0.00670 & 57.3 & 54.8 \\
power-law & 2 & 1 & 0.00461 & 60.3 & 59.0 & 0.00721 & 61.3 & 60.0 \\
exponential & 3 & 1 & .000814 & 15.2 & 14.0 & 0.00118 & 12.5 & 11.2 \\
logarithmic & 4 & 1 & .000635 & 8.75 & 7.49 & 0.00115 & 11.7 & 10.4 \\
Weibull & 5 & 2 & .000559 & 7.45 & 7.45 & 0.00136 & 18.2 & 18.2 \\
quadratic log&6 & 2 & {\bf .000460} & 2.40 & 2.40 & .000915 & 7.59 & 7.59 \\
Yule & 7 & 2 & .000788 & 16.4 & 16.4 & 0.00117 & 14.3 & 14.3 \\
Menzerath-Altmann/Inverse-Gamma
& 8 & 2 & 0.00251 & 46.5 & 46.5 & 0.00340 & 43.0 & 43.0 \\
Cocho/Beta & 9 & 2 & {\bf .000420} & 0 & 0 & {\bf .000691} & 0 & 0 \\
Frappat &10 & 3 & .000587 & 10.7 & 12.0 & .000838 & 7.20 & 8.49 \\
\hline
\end{tabular}
\end{center}
\caption{
\label{table3}
Regression diagnosis and model selection of ten functions
on English and Spanish letter rank-frequency plots.
}
\end{table}
Table \ref{table3} shows the AIC model selection result for the fitting
in Fig.\ref{fig3} and Fig.\ref{fig4}. The best function for both English
and Spanish, selected by either AIC or BIC, is the
Cocho/Beta function (Fig.\ref{fig3}(H), Fig.\ref{fig3}(H)). The second best
function is the quadratic logarithmic function (Fig.\ref{fig3}(E), Fig.\ref{fig4}(E)).
For English text, these functions are followed by Weibull,
logarithmic, and Frappat functions. For Spanish texts,
the two best functions are followed by Frappat,
logarithmic, and exponential functions.
A single SSE value does not tell us whether there exist
systematic deviations (e.g., larger deviations at high
rank numbers). To address this question, Fig.\ref{fig5}
and Fig.\ref{fig6} show the deviation at any rank number
for all fitting functions, for English and Spanish respectively.
It is interesting that functions with better fitting performance
all have a similar pattern in rank-specific deviation.
\section{Piecewise functions}
\indent
The zero-parameter Gusein-Zade function corresponds
to a simple exponential cumulative distribution (CD)
(for more discussions, see \citep{entropy10}):
\begin{equation}
CD= 1- \frac{r}{n+1} = 1- e^{-f/C}.
\end{equation}
In other words, the proportion of values that are larger than
$f_0$ is equal to $e^{-f_0/C}$.
Since Gusein-Zade function (Eq.(1)) can also be written as
$C= f/\log[ (n+1)/r]$, if we plot $f_i/\log( (n+1)/r_i)$
against $r_i$ ($i=1,2, \cdots n$), this function predicts
a plateau.
\begin{figure}[th!]
\begin{center}
\begin{turn}{-90}
\epsfig{file=fig7-step-function.eps, width=9cm}
\end{turn}
\end{center}
\caption{
\label{fig7}
An alternative form of Gusein-Zade function is
$f/\log(\frac{n+1}{r})= C$, and its validity can be
checked by plotting $f/\log(\frac{n+1}{r})$ versus the
rank $r$.
}
\end{figure}
Fig.\ref{fig7} shows $f_i/\log[ (n+1)/r_i]$ as a function
of rank, for both English (black) and Spanish (red) letters.
Surprisingly, instead of a plateau, we see step functions.
For English letters, the top 21 letters
({\sl etoniarshldcufmpwygbv}) form the first group,
and the next 5 letters ({\sl kxjqz}) form the second one.
The average plateau height of the first group in Fig.\ref{fig7} is
0.0425, that of the second group is 0.0157.
For Spanish letters in Fig.\ref{fig7}, three groups
appear in a step function. The two rarest letters
({\sl kw}) are very different from others (average plateau
height is 0.00165). This is a known fact as {\sl k}
and {\sl w} are only used in foreign words. The top 14 letters ({\sl eaosnirldctump})
are in one group (average height of 0.0437), and the
next 11 letters ({\sl bgyvfqhjzx\~{n}}) form the second
group (height is 0.0185). When the Spanish data is
compared to the English data, it is interesting that
the plateau height of the two groups are similar
across the language, whereas the number of letters
in the lower-plateau is much larger in Spanish than in English.
The result of Fig.\ref{fig7} indicates that we may
construct a piecewise Gusein-Zade function to fit
the ranked letter frequency distribution. It should
be noted that the number of parameters in a piecewise
Gusein-Zade function is no longer zero. For two-piece
function, three parameters are estimated:
plateau height of the first ($C_1$) and the second segment
($C_2 \ne C_1$),
and the partition position in $x$-axis ($r_0$). This minus
the normalization constraint leads to 2 free parameters.
This 3-parameter (2 of them are free) piecewise function
can be written as:
\begin{equation}
f = \left\{
\begin{array}{ll}
C_1 log \frac{n+1}{r} & \mbox{if $r < r_0$} \\
C_2 log \frac{n+1}{r} & \mbox{if $r \ge r_0$ } \\
\end{array}
\right.
\end{equation}
For the English letter data in Fig.\ref{fig3},
$r_0$ is chosen at 22, least square regression leads
to $C_1=0.04065$ and $C_2=0.01394$, and SSE= 0.000578.
For Spanish letters, with 2-segment function partition
at $r_0=15$, $C_1=0.0424$ and $C_2=0.01897$, and
SSE= 0.000539. Using 3-segment function, SSE is improved
only slightly to 0.000537. These results are comparable
to the best SSE results obtained by the Beta function
(Table \ref{table3}).
\section{Discussion}
\indent
So far we have not considered space as a ``letter". The number
of space is simply equal to the number of words ($N_{space}=N_{word}$),
and the space frequency is $p_{space}=N_{space}/(N_{space}+N_{letter})$.
For the US presidential speeches, the averaged $p_{space}$ is
0.174. For Mexican presidential speeches, the averaged $p_{space}$
is 0.162. There is a mild upward trend for $p_{space}$ in US
presidential speeches, but such a temporal pattern is missing
in Mexican texts.
When the ``space" is considered as a symbol, its frequency is
higher than any other single letters. The rank-frequency plot
with space symbol can still be fit perfectly by the Cocho/Beta
function (result not shown). The Cocho/Beta is still the best
function than others. However, the fitted coefficient values
can be quite different when space-symbol is included. For example,
for English texts, $a=0.21$ and $b=1.35$ without the space,
but $a=0.50$ and $b=0.875$ with the space symbol.
Due to the limited range of abscissa, many functions seem to fit the
ranked letter frequency distribution very well, and any subtle
change might disturb the relative performance among fitting functions.
Take the $-\log(r)$ type functions for example, we have considered
three similar functions already, Eq.(1), Eq.(4), Eq.(5), and Eq.(6).
The quadratic logarithmic function Eq.(6) clearly outperforms Eq.(1)
and Eq.(4), and competes with Cocho/Beta function to become the best
fitting function. We notice that in Gusein-Zade's original publication
\citep{gusein2}, he proposed a function of the form
$f= (1/r+1/(r+1) \cdots 1/n)/n$, which also appeared in \citep{gamow}
after a random division of unit length problem by John von Neumann.
That function can be approximated by a $-\log(r)$ function.
The piecewise plateaus revealed by Fig.\ref{fig7} seem
to partition alphabets into discrete groups. For English,
rare letters {\sl k,x,j,q,z} form their own group, with
frequencies much lower than expected by the $\log((n+1)/r$ function.
For Spanish, besides the well know letter group of {\sl k,w},
we found another group with letters {\sl b,g,y,v,f,q,h, j,z,x,\~{n}}.
The height of the first plateau is about twice that of the
second plateau, for both English and Spanish. One hypothesis
is that these lower-than-expected rare alphabets were originally
paired as one letter, then each ancestral letter was split into two
letters. Two such pairs can be imagined for English (discard {\sl z}),
and five pairs for Spanish (discard {\sl \~{n}}).
Of the ten functions used in this paper, some explicitly include
the number of letters, $n$, as part of the modeling, whereas
others do not. Those with $n$ include Gusein-Zade, Weibull,
and Cocho/Beta. For some linguistic data, the value of $n$
is fundamentally undecided, for example, the number of words
in a language. It is argued that word distribution should be
better modeled by ``large number of rare events" (LNRE)
\citep{baayen}. One consequence of LNRE is that the number
of words $n$ increases with the text length (followed the Heaps' law
\citep{heaps}), making the value of $n$ uncertain. Fortunately,
in letter frequencies, the value of $n$ is independent of the text length.
There might be deeper reasons why Cocho/Beta outperforms
nine other functions in fitting our data. It was
suggected that when a new random variable is constructed
by allowing both addition and subtraction of
independent and identically distributed random variables,
but within certain range, the new random variable follows
the Cocho/Beta distribution \citep{beltran}. Perhaps
Cocho/beta function is a limiting functional form for
ranked data under a very general condition.
In conclusion, we use ten functions to fit the English
and Spanish ranked letter frequency distribution obtained from
the US and Mexican presidential speeches. Cocho/Beta function
is the best fitting function among the ten, judged by sum of
errors (SSE) and Akaike information criterion (AIC). The
quadratic logarithmic function is a close second best.
We also discover a grouping of letters in both English and
Spanish. The rarer-than-expected group in English consists
of two pairs of letters whereas that in Spanish consists of
five pairs. There is a third, even-rarer-than-expected letter
group in Spanish with {\sl k,w}, consistent with the fact that
these are only used for foreign words. Besides the Cocho/Beta
and quadratic logarithmic function, it is not conclusive whether
other functions follow a universal relative fitting performance
order. Needless to say, studying letter frequencies in other languages
could potentially answer this question.
\section*{Appendix: Relationship between AIC/BIC and SSE}
\indent
Akaike information criterion is defined as: $AIC= - 2\log \hat{L} + 2p$,
where $\hat{L}$ is maximized likelihood, $p$ is the number of
parameter in the statistic model. When a dataset is fitted
by a model, if the error is normally distributed, the likelihood
of the model is ($n$ is the number of samples, $\sigma$ is the standard
deviation of the normal distribution for the error, $\{ y_i \}$
are the data points, and $\{ \hat{y_i} \}$ are the fitted value):
\begin{equation}
L = \prod_{i=1}^n \frac{e^{ -(y_i-\hat{y})^2/2\sigma^2} }{\sqrt{2\pi \sigma^2}}
=
\frac{ e^{ -\sum_i^n (y_i-\hat{y})^2)/2\sigma^2}}
{ (2\pi \sigma^2)^{n/2} }
\end{equation}
The $\sum (y_i-\hat{y_i})^2$ term can be called SSE (sum of squared errors).
If the error variance is unknown, it can be estimated from the data:
\begin{equation}
\hat{\sigma}^2= \frac{SSE}{n}
\end{equation}
Replacing $\sigma$ by the estimated $\hat{\sigma}$, we obtained
the maximized likelihood, which after log is \citep{ripley}:
\begin{equation}
\log(\hat{L}) = C - \frac{n}{2} \log(\hat{\sigma}^2)
= C - \frac{n}{2} \log(SSE/n)
\end{equation}
then,
\begin{equation}
AIC = n \log(SSE/n) + 2 \cdot p + const.
\end{equation}
and
\begin{equation}
BIC = n \log(SSE/n) + \log(n) \cdot p + const.
\end{equation}
\section*{Acknowledgements}
We would like to thank Osman Tuna G\"{o}kg\"{o}z for introducing
us the work by Al-Kindi. This work was partially supported by
UNAM-PAPIIT project IN115908. The authors wish to thank the
hospitality of the Centro de Investigaci\'{o}n en Matem\'{a}ticas
Aplicadas, Pachuca, M\'{e}xico, where the draft was finalized.
|
\section{Introduction}
A supernova remnant (SNR) consists of expelled material called
``ejecta'' from the explosion and a swept-up interstellar matter.
The X-ray emission results from the interactions of ejecta and
swept-up matter. From the X-ray emission of SNRs we may obtain
valuable information about the physical properties of the ejecta,
swept-up plasma, elemental abundances and the history of the
explosion. Shock wave may be the blast wave associated with the
stellar explosion and/or the reverse shock wave, which propagates
inwards from the decelerated blast wave and raises the temperature
of the stellar ejecta. The young SNRs are bright in X-rays and
dominated by the emission from ejecta. Thus they provide fruitful
information about the elements synthesized by the supernova (SN)
explosions. Therefore, the observation of the young SNRs is the
best method to investigate the abundances of the elements
synthesized by the SNe (see e.g. SN1006 \citep {b11}, RCW86 \citep
{b3}, Tycho \citep {b24} for the best studied ejecta-dominated
SNRs in the X-ray band).
G346.6-0.2 ($\rmn{RA}(2000)=17^{\rmn{h}} 10^{\rmn{m}}
19^{\rmn{s}}$, $\rmn{Dec.}~(2000)=-40\degr 11\arcmin$) is a
shell-type SNR located in the Galactic plane. It was discovered by
\citet {b5} in radio band having an angular size of 8 arcmin
\citep {b25}. In X-ray band, on the other hand, G346.6-0.2 was
first observed by {\it ASCA} during its Galactic plane survey
\citep {b28}. It is shown that the size of the X-ray emission from
G346.6-0.2 is less extended than its reported radio structure.
Five OH(1720 MHz) masers were detected toward this SNR and they
are all located along the southern edge of the remnant \citep
{b10}.
We are studying the X-ray emissions from the ejecta-dominated
SNRs. For this purpose we have chosen several small size Galactic
SNRs, one of which is G346.6-0.2. Detailed properties of
G346.6-0.2 remained unknown so far because of very limited photon
statistics in the {\it ASCA} ({\it AGPS}) data. {\it Suzaku} is
the most recent X-ray astronomical satellite (see \citet {b17})
having a large collecting area and low background. Therefore, it
is the best instrument for observing dim and diffuse sources. We
propose here to study G346.6-0.2 to understand the origin of its
thermal and non-thermal X-ray emission which will help us to
distinguish its SN explosion type. By using the archival data of
{\it Suzaku}, we were able to produce higher quality image and the
spectra of the remnant, which lead to the results in this study.
In Section 2, we describe the observation log and the data
reduction methods. We present the image analysis in Section 2.1
and the spectral analysis in Section 2.2. We discuss our results
and the origin of the thermal and non-thermal emission in Section
3.
\section[]{Observation and Data Reduction}
{\it Suzaku} satellite has two sets of instruments, one being the
four X-ray imaging spectrometers (XISs, see \citet {b12}); each of
four are at the focus of an X-ray telescope (XRTs, see \citet
{b21}) and also a separate hard X-ray detector (HXD, see \citet
{b22}; \citet {b9}). XIS's have two different type of CCDs: one
being three front illuminated CCDs (namely, FI, XIS0, XIS2 and
XIS3) and other is a back illuminated CCD (namely, BI, XIS1). Each
CCD covers an area of 17.8$\times$17.8 arcmin$^{2}$.
The data used in this analysis is taken by the XIS onboard {\it
Suzaku}. G346.6-0.2 was observed on 2009 October 07 for 56.7 ks
(Obs ID:504096010). During the observation, the XIS were operated
in the normal full-frame clocking mode with the editing mode of
$3\times3$ and $5\times5$, which are low and medium data rates,
high and super-high data rates, respectively.
For the data reduction and spectral analysis, we used the {\sc
headas} software package of version 6.5 and {\sc xspec} version
11.3.2 (see \citet {b2}). The response matrix files (RMF) and
ancillary response files (ARF) were made using {\sc xisrmfgen} and
{\sc xissimarfgen} version 2006-10-17 (see \citet {b7}).
\subsection{Image Analysis}
Figure 1 shows the XIS0 image of G346.6-0.2 in the $0.3-10$ keV
energy band. The solid circle (radius $\sim$5 arcmin) from the
centre) shows the region where the spectra is extracted. The
background region is shown by dotted circle
($\rmn{RA}(2000)=17^{\rmn{h}} 10^{\rmn{m}} 54^{\rmn{s}}$,
$\rmn{Dec.}~(2000)=-40\degr 06\arcmin 24\arcsec$, radius $\sim$1.8
arcmin). We excluded the upper right corner of the detectors
containing the onboard masked out $^{55}$Fe calibration sources.
For comparison radio continuum image of G346.6-0.2 at 843 MHz by
\citet{b25} is overlaid on the figure, where four maser sources
are also pointed out as has been reported by \citet {b10}.
The bright image in XIS0 is clearly seen and we have obtained the
Si-S map in $1.5-2.5$ keV energy band (Figure 2) to see if this
emission is from interstellar medium (ISM) or from the ejecta. In
Fig. 2, Si-S emission is distributed throughout the remnant and it
is bright, which may well indicate that the X-ray emission coming
from the remnant is ejecta dominated rather than originating from
the ISM.
\subsection{Spectral Analysis}
The extracted XIS spectra from a circular region of G346.6-0.2
(see Fig. 1) are analyzed in the full energy band of $0.3-10$ keV.
Figure 3 shows representative spectra of XIS0, XIS1 and XIS3
simultaneously, extracted from the region (shown in Fig. 1 by the
solid circle) with its corresponding best-fitting model and
residuals.
The strong Si and S lines from the spectra and Si-S image (see
Fig. 2) in $1.5-2.5$ keV energy band suggest that the emission is
ejecta dominated. Therefore, we have applied one-component models
in our spectral fits. We first fitted the data with an absorbed
(wabs; \citet {b18}) VNEI, which is a model for a non-equilibrium
ionization (NEI) collisional plasma with variable abundances
\citep {b4}. The model did not fit well (reduced $\chi^{2}$ of
1.12 (637/567 d.o.f.)). So, we added a power-law model
(non-thermal synchrotron emission from relativistic electrons).
The parameters of the absorbing column density ($N_{\rm H}$);
electron temperature ($kT_{\rm e}$); the ionization parameter
($n_{\rm e}t$) and the metal abundances C, N, O, Ne, Mg, Si, S,
Ar, Ca, and Fe are set free in this fit resulting in an
unacceptable error value. Then, we set Mg, S, Si, Ca, and Fe free,
since their emission lines are clearly seen in the spectra, while
the rest are set to solar values \citep{b1} resulting in a
reasonably good reduced $\chi^{2}$ value of 1.07 (606/565 d.o.f.)
with reasonable error limits. Similar steps were also applied to
VMEKAL model, which describes an emission spectrum from hot
ionized gas in collisional ionization equilibrium with variable
abundances \citep {b15, b16, b13}, and VPSHOCK model, which is
suitable for modelling plane-parallel shocks in young SNRs, where
plasma has not reached the ionization equilibrium \citep {b4}.
These two models gave us reduced $\chi^{2}$ value of 1.22 (691/568
d.o.f.) and 1.11 (629/567 d.o.f.), respectively. Then we added a
power-law component to both models, which yielded almost similar
results (of reduced $\chi^{2}/$d.o.f. value of 607/566 and
605/565, respectively, of the best-fitting parameters and errors
values) as VNEI model as given in Table 1.
These three models basically represent the emission from optically
thin thermal plasma with slight differences. For example, VNEI is
a NEI model similar to VPSHOCK, except with a single ionization
timescale.
From an absorbed VNEI and power-law model, the absorbing column
density, the electron temperature, the ionization parameter, the
metal abundances of Mg, Si, S, Ca, Fe, volume emission measure
(VEM) and flux ($F_{\rm x}$) of the VNEI component and the photon
index ($\Gamma$) and norm of power-law component are obtained and
given in Table 1, with the corresponding error values with 90 per
cent confidence levels. Total flux and reduced $\chi^{2}/$d.o.f.
value of VNEI and power-law components are also given in this
table. The results of VMEKAL and power-law, VPSHOCK and power-law
models are also presented in Table 1 for comparison.
To determined the line centre energy of the K-shell ($K\alpha$)
lines, we fitted the spectra using a simple bremsstrahlung
continuum plus five Gaussian lines with an absorbing column. The
obtained best-fitting central energies of $K\alpha$ emission lines
are He-like Mg (1.38$\pm0.06$ keV), Si (1.86$\pm0.02$ keV), S
(2.52$\pm0.06$ keV), Ca (3.25$\pm0.55$ keV) and Fe (6.77$\pm0.32$
keV).
Figure 4 shows the best-fitting metal abundances (normalized to
Si) relative to solar values \citep{b1} with the predicted
nucleosynthesis W7 Type Ia SN model \citep{b19}, which is widely
used as the standard model.
\section{Discussion and Conclusions}
In this work, we provide for the first time a high-quality image,
spectra, and detailed description of the X-ray emission of
G346.6-0.2 using a set of public archival {\it Suzaku} data. The
radio emission of this remnant extends to 8 arcmin \citep {b25},
while X-ray image shows a smaller extension (see Fig. 1), which is
consistent with that of the $ASCA$ results shown in \citet {b28}.
We fitted the spectra with ``thermal'' component in
non-equilibrium ionization and an additional ``non-thermal''
power-law component with reasonably good chi-square values.
\emph{Thermal emission.} The thermal part of the spectra can be
represented by non-equilibrium ionization (VNEI) model. The
thermal X-ray emission of young SNRs, whose emission is still
dominated by the ejecta, is predominantly from the ejecta heated
by the reverse shock. Since the ejecta is metal abundant, its
X-ray spectra usually shows strong emission lines of heavy
elements such as O, Ne, Mg, Si, S, Ar, Ca and Fe. $K\alpha$
emission lines of Mg, Si, S, Ca and Fe for G346.6-0.2 are detected
clearly for the first time in this work.
The interstellar absorbing column density $N_{\rm H}$ is obtained
to be ($2.1\pm 0.2$)$\times10^{22}$ $\rm cm^{-2}$, which is
consistent with the galactic value of ($1.5\pm
0.2$)$\times10^{22}$ $\rm cm^{-2}$ in that direction \citep {b6}.
$ASCA$ results of G346.6-0.2 reported a value of $N_{\rm
H}>10^{22}$ $\rm cm^{-2}$ \citep {b28}, indicating that X-ray
spectra of the remnant is heavily absorbed by interstellar matter,
perhaps due to its location being close to the Galactic plane. In
other words, the electron density ($n_{\rm e}$) of medium of the
remnant is high enough. This may also be supported by the fact
that the south region of the remnant is in interaction with four
molecular clouds as has been noted by \citet {b10}.
Being close to the Galactic plane and being in a highly dense
region, one can expect that some portions of the remnant can be
highly asymmetrical. The asymmetry in the south region of the
remnant may well be due to the interaction with the maser sources
as has been noted in Fig. 1. One can argue that the surface
brightness of the remnant is far from the equal distribution and
spherical symmetry due to the inhomogeneity of the medium.
The thermal emission further provides us information about the
electron temperature and age of the plasma. Our best-fitting
temperature is obtained to be $\sim$1.2 keV, which is typical for
shell-like SNRs. This value is also consistent with the result of
$\sim$1.6 keV obtained from {\it ASCA} observations \citep {b28}.
The age of G346.6-0.2 has not been predicted so far. Especially,
young SNRs whose ages are a few hundred or about a thousand years
are not only bright but also dominated by the emission from
ejecta. Being an ejecta-dominant remnant and bright in X-rays
G346.6-0.2 can be a young SNR. Plasma in most of young SNRs are in
the NEI condition. The ionization age (or ionization parameter) is
defined as $\tau$=$n_{\rm e}t$, which is often used as a key
diagnostic of the NEI state. $\tau$ is typically required to be
$\geq$$10^{12}$ $\rm cm^{-3}$s \citep {b14} for full ionization
equilibrium. In our observations we obtained $n_{\rm e}t$ to be
(2.92$\pm0.01$)$\times10^{11}$ $\rm cm^{-3}$s, which may also
indicate that this SNR is far from full ionization equilibrium.
So, the plasma has not yet had time to reach ionization
equilibrium, and it is still being ionized. We calculated $n_{\rm
e}$ to be $\sim$0.82 $\rm cm^{-3}$ from VEM=$n_{\rm e}n_{\rm H}V$
where $n_{\rm e}=1.2n_{\rm H}$ and V is the X-ray-emitting volume
of the remnant (estimated to be $\sim$$2\times10^{54}$ $\rm
cm^{3}$) by adopting an average distances of d $\sim8.3$ kpc
(average of a near value of 5.5 kpc and a far value of 11 kpc
given by \citet {b10}) and assuming the emitting volume to be
spherical shell. From the equation t=$\tau$/$n_{\rm e}$, the age
of G346.6-0.2 calculated to be $\sim$$1.1\times10^{4}$ yr. Thus,
this remnant is likely to be the oldest known ejecta-dominated
shell-like SNR (e.g. SN1006, RCW86, G337.2-0.7 and G309.2-0.6,
$\sim$1000 yr \citep {b26}, $\sim$1800 yr \citep {b30}, 2000-4500
yr and 700-4000 yr \citep {b20}, respectively.)
\emph{Non-thermal emission.} Recently, in some SNRs (e.g. first
SN1006 \citep {b11} and then RX J1713.7.3946 \citep {b29, b34},
Cas A \citep {b32, b33}, and Tycho's SNR \citep {b31}) non-thermal
emission has been detected. There are two possible mechanisms for
non-thermal emission: one is synchrotron emission and the another
is non-thermal bremsstrahlung. The non-thermal emission coming
from G346.6-0.2 is most likely a synchrotron emission (power-law
component with photon index of $\Gamma$ $\sim 0.6$). The real
source of synchrotron emission in SNRs is believed to be the
acceleration of the ultrarelativistic electrons in the magnetic
fields with the help of the shock waves. A powerful X-ray point
source inside the remnant or in the vicinity of the remnant can
also contribute to the non-thermal emission. The photon index we
obtained ($\Gamma$ $\sim 0.6$) is lower (harder) than those of
typical shell-like SNRs. This forces us to presume that a pulsar
wind nebula (PWN) contribution is also possible, although a PWN
has not been discovered so far in or in the vicinity of
G346.6-0.2.
The flux value of G346.6-0.2 is obtained to be $F_{\rm x}$
$\sim4.4\times10^{-11}$ erg $\rm s^{-1}\rm cm^{-2}$ for $0.3-10$
keV range, which is in a good agrement with that of found with
$ASCA$ \citep {b28}. Using this flux value for the source and also
taking its X-ray angular size of $\sim$5 arcmin, its surface
brightness in the $0.3-10$ keV energy range is found to be
$\Sigma$ $\sim1.8\times10^{-12}$ erg $\rm s^{-1}\rm cm^{-2}\rm
arcmin^{-2}$, which is higher than many shell-like Galactic SNRs.
Surface brightness mainly depends on the explosion energy and the
density of the medium, but a PWN may also contribute to the
surface brightness of the SNR. In this case, high medium density
and small size of the remnant may be the main reason of high X-ray
surface brightness of G346.6-0.2.
\emph{Relative abundances in the ejecta. } Theoretical models
predict that in core collapse SN explosion low-Z elements like O,
Ne, Mg are produced \citep {b23}, while in Type Ia SNe \citep
{b19} high-Z elements like Ar, Ca and Fe are mostly produced.
Furthermore, Fe production in Type Ia SNe is far larger than that
of core-collapse SNe \citep {b19,b8}. The detection of Fe emission
lines from the remnants is important to identify the type of the
remnant. $Suzaku$ has detected Fe $K\alpha$ lines in especially
Type Ia SNRs (e.g. Tycho, SN1006, RCW86).
When we compare our best-fitting abundances of Mg, S, Ca and Fe
relative to Si with the predicted theoretical values, Mg relative
to Si is higher, while S and Ca relative to Si are lower. The
reason for the lower abundances of S and Ca could be because the
elements are concentrated at the inner layers of the remnant and
hence are not heated enough by the reverse shock yet. This
information implies that this SNR is at its early evolution phase.
Our best-fitting abundance Fe relative to Si is consistent with
the expected value (see Fig. 4) within the confidence range.
However, considering the large error bars due to low statistics,
we could also comment a similar statement for Fe abundance value
and an unattained reverse shock, as it is reported in many Type Ia
SNRs like SN1006 \citep {b26}, Tycho \citep {b27} and G337.2-0.7
\citep {b20} the Fe abundance is being less than expected value
produced in a Type Ia SNe. Considering all these cases we can
predict that the remnant may be originated from a Type Ia SN
explosion.
\section*{Acknowledgments}
A.S. is supported by T\"{U}B\.{I}TAK Post$-$Doctoral Fellowship.
This work is supported by the Akdeniz University Scientific
Research Project Management and by T\"{U}B\.{I}TAK under project
codes 108T226 and 109T092. The authors also acknowledge the
support by Bo\u{g}azi\c{c}i University Research Foundation under
2010-Scientific Research Project Support (BAP) project no:5052.
|
\section{Introduction\label{sec:Introduction}}
The \emph{Fr\'{e}chet distance} is a similarity metric for continuous shapes such as curves and surfaces. In the case of computing it between two (directed open) curves there is an intuitive explanation of the Fr\'{e}chet distance. Suppose a man walks along one curve, a dog walks along the other, and they are connected by a leash. They can vary their relative speeds but cannot move backwards. Such a walk pairs every point on one curve to one and only one point on the other curve (i.e., creates homeomorphism between surfaces) in a continuous way. The Fr\'{e}chet distance\ of the curves is the minimum leash length required for the man and dog to walk along these curves. Although less intuitive, the idea is similar for surfaces.
While the Fr\'{e}chet distance\ between polygonal curves can be computed in polynomial time \cite{Alt1995}, computing it between surfaces is much harder. In \cite{godau} it was shown that even computing the Fr\'{e}chet distance\ between a triangle and a self-intersecting surface is NP-hard. This result was extended in \cite{bbs-fds-10} to show that computing the Fr\'{e}chet distance\ between 2d terrains as well as between polygons with holes is also NP-hard. Furthermore, while in \cite{Alt2010} it was shown to be upper semi-computable, it remains an open question whether the Fr\'{e}chet distance\ between general surfaces is even computable.
On the other hand, in \cite{Buchin2006} a polynomial time algorithm is given for computing the Fr\'{e}chet distance\ between two (flat) simple polygons. This was the first paper to give any algorithm for computing the Fr\'{e}chet distance\ for a nontrivial class of surfaces and remains the only known approach for computing it. The main idea of their algorithm is to restrict the kinds of different mappings that need to be considered. Our contribution is to generalize their algorithm to a class of non-flat surfaces we refer to as folded polygons. Given that theirs is the only known approach it is of particular importance to explore extending it to new classes of surfaces. The major problem we encountered in generalizing the work of \cite{Buchin2006} was that the kinds of different of mappings which need to be considered is less restricted. We address three different methods to resolve this problem. In Section \ref{sec:FPTAlgorithm}, we outline a fixed-parameter tractable algorithm. In Section \ref{sec:GeneralApprox}, we describe a polynomial-time approximation algorithm to compute the Fr\'{e}chet distance\ between folded polygons\ within a constant factor. In Section \ref{sec:LinfParallel}, we describe a nontrivial class of folded polygons\ for which the original algorithm presented in \cite{Buchin2006} will compute an exact result.
\vspace{-3 mm}
\section{Preliminaries\label{sec:Preliminaries}}
The Fr\'{e}chet distance\ is defined for two k-dimensional hypersurfaces $P,Q:[0,1]^{x}\rightarrow\mathbb{R}^{d}$, where $x \leq d$,
as \[
\delta_{F}(P,Q)\ =\ \inf_{\sigma:A\rightarrow B}\ \sup_{p\in A}\ \|P(p)-Q(\sigma(p))\|\]
where $\sigma$ ranges over orientation-preserving homeomorphisms
that map each point $p\in P$ to an image point $q=\sigma(p)\in Q$. Lastly, $\|\cdot\|$ is the Euclidean norm but other metrics could be used instead.
Let $P,Q:[0,1]^{2}\rightarrow\mathbb{R}^{d}$ be connected polyhedral surfaces for each of which we have a convex subdivision. We assume that the dual graphs of the convex subdivisions are acyclic, which means that the subdivisions do not have any interior vertices. We will refer to surfaces of this type as \emph{folded polygons}. We refer to the interior convex subdivision edges of $P$ and $Q$ as \emph{diagonals} and \emph{edges} respectively. Let $m$ and $n$ be the complexities of $P$ and $Q$ respectively. Let $k$ and $l$ be the number of diagonals and edges respectively. Assume without loss of generality the number of diagonals is smaller than the number of edges. Let $T_{matrixmult}(N)$ denote the time to multiply two $N\times N$ matrices.
\vspace{-3 mm}
\subsection{Simple Polygons Algorithm Summary \label{sec:SimpleSummary}}
In previous work Buchin et al$.$ \cite{Buchin2006} compute the
Fr\'{e}chet distance\ between simple polygons $P$, $Q$. The authors show that, while the Fr\'{e}chet distance\ between a convex polygon and a simple polygon is the Fr\'{e}chet distance\ of their boundaries, this is not the case for two simple polygons. The idea of their algorithm is to find a convex subdivision of $P$ and map each of the convex regions of it continuously to distinct parts of $Q$ such that taken together the images account for all of $Q$.
First, they show that the decision
problem $\delta_{F}(P,Q)\leq\varepsilon$ can be solved by (1) mapping
the boundary of $P$, which we denote by $\partial P$, onto the
boundary of $Q$, which we denote by $\partial Q$, such that $\delta_{F}(\partial P,\partial Q)\leq\varepsilon$
and (2) mapping each diagonal $d$ in the convex subdivision of $P$
to a shortest path $f \subseteq Q$ such that both endpoints of $f$ lie
on $\partial Q$ and such that $\delta_{F}(d,f)\leq\varepsilon$.
In order to solve subproblem (1) they use the notion of a free space diagram. For open curves $f,g:[0,1]\rightarrow\mathbb{R}^{d}$ it is defined as $FS_{\varepsilon}(f,g) = \{(x,y)\ |\ x \in f,\ y \in g,\ ||x-y||\leq\varepsilon\}$ where $\varepsilon\ \geq 0$. A monotone path starting at the bottom left corner of the free space diagram\ going to the top right exists if and only if the curves are within Fr\'{e}chet distance\ $\varepsilon$. As shown in \cite{Alt1995}, this can be extended to closed curves by concatenating two copies of the free space diagram\ to create a \emph{double free space diagram}\ and searching for a monotone path which covers every point in $P$ exactly once see Figure \ref{fig:FSDpathNew}.
This algorithm can be used to show whether $\delta_{F}(\partial P,\partial Q)\leq\varepsilon$ and find the particular mapping(s) between $\partial P$ and $\partial Q$. In turn this defines a \emph{placement} of the diagonals, i.e., a mapping of the endpoints of the diagonals to endpoints of the corresponding image curves in $Q$.
Subproblem (2) is solved by only considering paths through the free space diagram\ that map a diagonal $d$ onto an image curve $f$ such that $\delta_{F}(d,f)\leq\varepsilon$. Naturally the particular placement of the diagonals determined in subproblem (1) could affect whether this is true. Therefore, they must check this for many paths in the free space diagram. Fortunately they can show that it is sufficient to only consider mapping a diagonal to an image curve which is the shortest path between the end points determined by the placement.
Solving these subproblems generates a mapping between $P$ and $Q$ for $\varepsilon$. This mapping might not be a homeomorphism but the authors show by making very small perturbations of image curves the mapping can be made into one. These perturbations can be arbitrarily small. Thus, because the Fr\'{e}chet distance\ is the infimum of all homeomorphisms on the surfaces, the Fr\'{e}chet distance\ is $\varepsilon$. For simplicity we will refer to these generated mappings as homeomorphisms. By performing a binary search on a set of critical values they can use the above algorithm for the decision problem to compute the Fr\'{e}chet distance\ of $P$ and $Q$.
\vspace{-3 mm}
\begin{figure*}[ht]
\centering \includegraphics[width=0.85\textwidth]{DoubleFSExample-01}
\centering \caption{\small The white areas are those in free space diagram. The surfaces are within Fr\'{e}chet distance\ $\varepsilon$ since there is a monotone path starting at the bottom of the free space diagram\ and ending at the top which maps every point on the boundary of $P$ exactly once. This figure was generated using an ipelet created by G\"{u}nter Rote.
\label{fig:FSDpathNew}}
\vspace{-3 mm}
\end{figure*}
\subsection{Shortest Path Edge Sequences \label{sec:ShortestPathEdgeSequences}}
Our algorithm extends the simple polygons algorithm to one for folded polygons. The idea of the algorithm is to subdivide one surface, $P$, into convex regions and pair those with corresponding regions in the other surface, $Q$. The difference is that those regions of $Q$ are now folded polygons\ rather than just simple polygons. The authors of \cite{Buchin2006} show that the Fr\'{e}chet distance\ of a convex polygon and a simple polygon is just the Fr\'{e}chet distance\ of their boundaries. Using almost the same argument we prove that it also holds for folded polygons, see Appendix \ref{sec:ConvexVSFoldedPolygonProof} for the full proof
\begin{lemma}
\label{lem:GeneralApprox-ConvexVSFoldedPolygon}
The Fr\'{e}chet distance\ between a folded polygon\ $P$ and a convex polygon $Q$ is the same as that between their boundary curves.
\end{lemma}
As mentioned in Section \ref{sec:SimpleSummary}, for the simple polygon algorithm it suffices to map diagonals onto shortest paths between two points on $\partial Q$. By contrast, there are folded polygons\ where a homeomorphism between the surfaces does not exist when diagonals are mapped to shortest paths but does exist when the paths are not restricted, see Figure \ref{fig:FSDpath}. The curve $s_1$ is the shortest path between the points $a$ and $b$ but the curve $s_2$ has smaller Fr\'{e}chet distance\ to $d$ than $s_1$ has. We must therefore consider mapping the diagonals to more general paths. Fortunately, we can show that these more general paths still have some nice properties for folded polygons.
\begin{figure*}[ht]
\vspace{-4 mm}
\centering \includegraphics[width=0.80\textwidth]{ShortestPathCE-01}
\centering \caption{\small An example where the image curve $s_2$ with the smallest Fr\'{e}chet distance\ to a diagonal $d$ is a non-shortest path $s_1$ in $Q$. (a) overhead view, (b) sideview.
\label{fig:SPCounterExample}\label{fig:FSDpath}}
\vspace{-3 mm}
\end{figure*}
\vspace{-3 mm}
\begin{lemma}
\label{lem:EdgePath}
Let $Q$ be a folded polygon, $u$ and $v$ be points such that $u,v \in \partial Q$, $E = \lbrace e_0, e_1, \ldots, e_s \rbrace$ be a sequence of edges in the convex subdivision of $Q$, and $d$ be a line segment. Given $\varepsilon > 0$ we can find a curve $f$ in $Q$ that follows the edge sequence $E$ from $u$ to $v$ such that $\delta_F (d,f) \leq \varepsilon$, if such a curve exists, in $O(s)$ time.
\end{lemma}
\proof
We construct a series $F = FS_{\varepsilon}(e_0,d), FS_{\varepsilon}(e_1,d), \ldots, FS_{\varepsilon}(e_s,d)$, of 2-dimensional free space diagrams. Any two edges $e_i$ and $e_{i+1}$ are on the boundary of the same convex polygon in the convex subdivision of $Q$ so we can assume without loss of generality that $f$ consists of straight line segments between the edge intersections. This is similar to the shortcutting argument used to prove Lemma 3 in \cite{Buchin2006}. Thus, we only need to check the points where $f$ crosses an edge of $Q$. For $\delta_F (d,f) \leq \varepsilon$ to be true, the preimages of those crossing points must be monotone along $d$. Let $FS_{\varepsilon}^{'}(e_i,d)$ be the projection of $FS_{\varepsilon}(e_i,d)$ onto $d$. Let $F^{'} = FS_{\varepsilon}^{'}(e_1,d), FS_{\varepsilon}^{'}(e_2,d), \ldots, FS_{\varepsilon}^{'}(e_s,d)$.
To verify that the preimage points on $d$ can be chosen such that they are monotone, we check the intervals of $F^{'}$. Specifically, for $i<j$, the point on $d$ mapped to $e_i$ must come before the one mapped to $e_j$. This can be checked by greedily scanning left to right and always choosing the smallest point on $d$ which can be mapped to some edge. A search of this form takes $O(s)$ time.
\qed
The dual graph of the faces of $Q$ is acyclic. This implies that there is a unique sequence of faces through which a shortest path from $u$ to $v$, where $u,v \in \partial Q$, must pass. Necessarily, there must also be a unique edge sequence that the shortest path follows. We call such an edge sequence the \emph{shortest path edge sequence}.
\begin{lemma}
\label{lem:OnlyShortestPathEdgeSequenceNeeded}
Let $d$ be a diagonal. If there is a curve $f \subseteq Q$ with $\delta_F (d,f) \leq \varepsilon$ then there is a curve $g \subseteq Q$ which follows the \emph{shortest path edge sequence} such that $\delta_F (d,g) \leq \varepsilon$.
\end{lemma}
\proof
Let $E_f$ and $E_g$ be the edge sequences of $f$ and $g$ respectively. By definition the dual graph of the faces of $Q$ is acyclic, so $E_g$ must be a subsequence of $E_f$. $E_g$ induces a sequence of free space intervals. If there is a monotone path in the free space interval sequence induced by $E_f$, we can cut out some intervals and have a monotone path in the free space for $E_g$.
\qed
From Lemma \ref{lem:OnlyShortestPathEdgeSequenceNeeded}, we just need to consider paths that follow the shortest path edge sequence. We refer to paths that follow this edge sequence and consist of straight line segments between edges as \emph{Fr\'{e}chet shortest paths}. In addition, $s$ in Lemma \ref{lem:EdgePath} is bounded by the number of edges in along the shortest path edge sequence between $u$ and $v$, and $E$ will be the shortest path sequence. This implies the following theorem.
\begin{theorem}
\label{thm:FrechetPathWithPoints}
Let $Q$ be a folded polygon, $u$ and $v$ be points such that $u,v \in \partial Q$, and $d$ be a line segment. Given $\varepsilon > 0$, we can in $O(l)$ time find a curve $f$ in $Q$ from $u$ to $v$ such that $\delta_F (d,f) \leq \varepsilon$ if such a curve exists.
\end{theorem}
Suppose we have a homeomorphism between $\partial P$ and $\partial Q$. The endpoints of the image curves must appear on $\partial Q$ in the same order as their respective diagonal endpoints on $\partial P$. The homeomorphism also induces a direction on the diagonals in $P$ and on the edges in $Q$. Specifically, we consider diagonals and edges to start at their first endpoint along $\partial P$ or $\partial Q$ respectively in a counterclockwise traversal of the boundaries. We denote by $D_{e}$ the set of diagonals whose associated shortest
path edge sequences contain an edge $e \subseteq Q$. Observe that pairwise
non-crossing image curves must intersect an edge $e$
in the same order as their endpoints occur on $\partial Q$. We refer
to this as the \emph{proper intersection order} for an edge $e$.
\vspace{-3 mm}
\subsection{Diagonal Monotonicity Test and Untangleability \label{sec:DiagonalConsistency}}
We now define a test between two folded polygons\ $P$ and $Q$ which we call the \emph{diagonal monotonicity test}. For a given $\varepsilon$ this test returns true if the following two things are true. First, $\delta_F(\partial P, \partial Q) \leq \varepsilon$. Second, for every diagonal $d_i$ in the convex subdivision of $P$, the corresponding Fr\'{e}chet shortest path\ $f_i$ in $Q$ has $\delta_F(d_i, f_i) \leq \varepsilon$. We refer to the class of mappings of the folded polygons\ generated by this test as \emph{monotone diagonal mappings}. This is similar to the test used by \cite{Buchin2006} except ours uses Fr\'{e}chet shortest paths\ instead of the shortest paths.
Unfortunately, because the image curves of the diagonals are no longer shortest paths, they may cross each other and we will no longer be able to generate a mapping between the folded polygons\ which is a homeomorphism. For an example of this see Appendix \ref{sec:CounterExample}. Thus the diagonal monotonicity test\ might return true when in fact a homeomorphism does not exist. We must explicitly ensure that the image curves of all diagonals are non-crossing. In particular, we refer to a set of image curves $F = \lbrace f_1 \ldots f_k \rbrace$ as \emph{untangleable} for $\varepsilon$ if and only if there exists a set of image curves $F^{'} = \lbrace f^{'}_1 \ldots f^{'}_k \rbrace$ where $f_i$ and $f^{'}_i$ have the same end points on $\partial Q$, $\delta_F (d_i,f^{'}_i) \leq \varepsilon$, and the curves of $F^{'}$ are pairwise non-crossing. A homeomorphism exists between the folded polygons\ for $\varepsilon$ if and only if there exists a monotone diagonal mapping\ whose image curves are untangleable for $\varepsilon$. This follows from the same argument used in \cite{Buchin2006} for simple polygons.
As shown in Theorem \ref{thm:FrechetPathWithPoints}, computing Fr\'{e}chet shortest paths\ instead of shortest paths does not increase the asymptotic run time. To optimize this $\varepsilon$ we can perform a binary search on a set of critical values. As in \cite{Buchin2006}, the number of critical values is $O(m^2 n + m n^2 )$. The three types of critical values between a diagonal and its corresponding path through $Q$ are very similar to those outlined in the simple polygons algorithm. So, by following the paradigm set forth by \cite{Buchin2006}, we arrive at the following theorem:
\begin{theorem}
\label{thm:SimpOptimized-Runtime}
The minimum $\varepsilon$ for which two folded polygons\, $P$ and $Q$, pass the diagonal monotonicity test\ can be computed in time $O(kT_{matrixmult}(mn) \log(mn))$.
\end{theorem}
\vspace{-3 mm}
\section{Fixed-Parameter Tractable Algorithm\label{sec:FPTAlgorithm}}
In this section we outline an algorithm to decide for a fixed mapping between the boundaries of a pair of folded polygons\ whether the image curves induced from the mapping are untangleable. From this we create a fixed-parameter tractable algorithm for computing the Fr\'{e}chet distance\ between a pair of folded polygons.
\vspace{-3 mm}
\subsection{Untangleability Space\label{sec:UDiagram}}
Let $e$ be an edge in $Q$ which is crossed by the image curves of $h$ diagonals, $d_1, \ldots, d_h$. We assume without loss of generality that the image curves of the diagonals cross $e$ in proper intersection order if, for all $1 \leq i,j \leq h$ where $i<j$, the image curve of $d_i$ crosses $e$ before the image curve of $d_j$ crosses $e$. Let the {\em untangleability space} $U_{e}$ contain all $k$-tuples of points on the diagonals which can be mapped to crossing points on the edge $e$ within distance $\varepsilon$ and such that the crossing points are in the proper intersection order along $e$. $U_{e}$ can be shown to be convex yielding the following theorem.
\begin{theorem}
\label{thm:ConvexUntangleabilitySpace}
$U_e(d_1, \ldots, d_k)$ is convex.
\end{theorem}
This theorem can be proven by linearly interpolating between points in $U_{e}$. For the full proof see Appendix \ref{sec:UntangleSpaceConvex}.
\vspace{-3 mm}
\subsection{Fixed-Parameter Tractable Algorithm\label{sec:Algorithm}}
\begin{minipage}{\textwidth}
We assume the complexity of $k$ and $l$, the convex subdivisions of $P$ and $Q$ are constant. Checking for the existence of a set of image curves which are untangleable can be done by using the untangleability spaces\ of the diagonals. Assume we are given some homeomorphism between $\partial P$ and $\partial Q$ from which we get a placement of the diagonals. We first choose an edge in $Q$ to act as the root of the {\em edge tree} that corresponds to the dual graph of $Q$. We propagate constraints imposed by each untangleability space\ up the tree to the root node to determine if the set of image curves induced by the placement of the diagonals is untangleable.
\begin{wrapfigure}{r}{0.15\textwidth}
\centering \includegraphics[width=0.14\textwidth]{UntangleabilityPropogation-01}
\centering \caption{}
\label{fig:Propogation-01}
\end{wrapfigure}
\hspace{4 mm} The untangleability space\ of an edge $e$, $U_{e}$, contains exactly those sets of points on the diagonals in $D_e$ which can be mapped to the edge in the proper intersection order. The point chosen in $U_{e}$ imposes a constraint on what points may be chosen in other untangleability spaces. In particular the corresponding points on all of the diagonals must be monotone with respect to their edge sequence. We define $C(U_{e_{i}})$ as the Minkowski sum of $U_{e_{i}}$ with a ray in the opposite direction of the constraint on each of the diagonals in $D_{e_{i}}$, see Figure \ref{fig:Propogation-01}. In Figure \ref{fig:Propogation-01} $U_e$ is shown in white and $C(U_e)$ is the union of the white and light gray portions. The direction of this constraint depends on which side of the edge $e_{i}$ the next edge is. $C(U_{e_{i}})$ contains exactly those sets of points on the diagonals not excluded from having a monotone mapping with $U_{e_{i}}$.
\end{minipage}
We define for every edge $e$ a $k$-dimensional {\em propagation space} $P_{e}$. If $e$ is a leaf in the tree, then $P_{e}=U_{e}$. Otherwise, define\[P_e=U_e \cap C(P_{e_1}) \cap \ldots \cap C(P_{e_j})\] where $e$ is the parent of the edges $e_{1}, e_{2}, \ldots, e_{j}$. $C(P_{e_{j}})$ contains only those points that are not excluded by the constraints of the tree rooted at $e_j$ from being used to untangle on the parent of $e_{j}$. The propagation space\ for the root will be empty if and only if this set of image curves are not untangleable. From our assumptions, the propagation space\ of the root can be computed in constant time as the intersection of semi-algebraic sets \cite{Basu2006}. Let $F(k,l)$ be the time complexity of computing this intersection.
Consider two different mappings between $\partial P$ and $\partial Q$. These determine different placements of the diagonals. If all of the image curves of all of the diagonals have the same shortest path edge sequence in both of the mappings the test will return the same result. Thus, we only need to test paths through the free space diagram\ which cross the diagonals and edges in a different order. The free space diagram\ for $\partial P$ and $\partial Q$ contains $2k$ vertical line segments that will each contribute $O(kl)$ different mappings of the diagonals and edges. Hence, there are $O((kl)^{2k})$ paths through the free space diagram\ which we need to test. For each of these we can check whether a global untangling exists as described above in constant time. Similar to the algorithm for polygonal curves \cite{Alt1995} we can perform Cole's \cite{Cole1987} technique for parametric search \cite{Megiddo1983} to optimize the value of $\varepsilon$. For convex subdivisions with constant number of edges $l$ and diagonals $k$, this yields a fixed-parameter tractable algorithm with runtime polynomial in $m$ and $n$.
\begin{theorem}
\label{thm:FixedParameter-Algorithm}
We can compute the Fr\'{e}chet distance\ of two folded polygons\ in time $O((F(k,l)(kl)^{2k} + kT_{matrixmult}(mn))\log(mn))$.
\end{theorem}
\vspace{-3 mm}
\section{Constant Factor Approximation Algorithm\label{sec:GeneralApprox}}
In this section we present an approximation algorithm based on our diagonal monotonicity test\ to avoid tangles altogether. First we demonstrate the following theorem:
\begin{theorem}
\label{thm:GeneralApprox-SimpleTest}
If two folded polygons, $P$ and $Q$, pass the diagonal monotonicity test\ for some $\varepsilon$, then $\delta_F(P, Q) \leq 9\varepsilon$.
\end{theorem}
\proof
Consider the image curves of the diagonals of $P$ found by performing our diagonal monotonicity test. To pass the diagonal monotonicity test\ there must be a homeomorphism between $\partial P$ and $\partial Q$. The image curves of the diagonals will be mapped in the proper order along the boundary of $Q$. Therefore, if a pair of image curves cross in $Q$ they must do so an even number of times.
Take two consecutive points $u$, $v$ on the boundary of $P$ that are connected by a diagonal $d= \overline{uv}$, and consider the convex ear of $P$ that $d$ ``cuts off''. Let $d^{'}$ be the image curve of $d$ in $Q$. In order to create a homeomorphism between $P$ and $Q$, $d^{'}$ should cut off an ear of $Q$ which can be mapped to the ear of $P$. Unfortunately, some image curves may cross this $d^{'}$ and cause tangles. Consider the arrangement of image curves in Q and let $d^{''}$ be the highest level of this arrangement closest to the top of the ear ($\partial Q$), such that $d^{''}$ connects the image points $u$ and $v$ on the boundary of $Q$.
\begin{figure}[ht]
\vspace{-6 mm}
\centering \includegraphics[width=0.9\textwidth]{FoldedApprox-02-combined02}
\centering \caption{\small
(a) the intersections between $d$ and $d_1$ imply that $\overline{ab}$ can be mapped to the region of $d_1^{'}$ between $a^{'}$ and
$b^{'}$ within Fr\'{e}chet distance\ $\varepsilon$
(b) this is the preimage of the intersection between $d_1^{'}$ and $d_2^{'}$ on $d$,
(c) an example of the preimages of two intersections occurring out of order of $d$}
\label{fig:FoldedApprox-02}
\vspace{-4 mm}
\end{figure}
Observe that if an image curve $d_1^{'}$ crosses $d^{'}$ from below then that intersection point $a^{'}$ has a pre-image on both $d$ and $d_1$. These points $a$ on $d$ and $a_1$ on $d_1$ can be no more than $2\varepsilon$ apart since they both map to the intersection within distance epsilon. In addition, $d_1^{'}$ must cross back below $d^{'}$ eventually since all image curves which cross do so an even number of times (take the first such occurrence after the initial crossing). The preimage points $b$ and $b_1$ of this second intersection $b$ are also no more than $2\varepsilon$ apart. Since both $d$ and $d^{'}$ are line segments, every point on the line segment $\overline{ab}$ on $d$ is $\leq 2 \varepsilon$ distance from some point on the line segment $\overline{a_{1}b_{1}}$ on $d_1$. For an approximation of $3\varepsilon$ we can map a point on $d$ to its corresponding point on $d_1$ and then to where that point maps on $d_1^{'}$, see Figure \ref{fig:FoldedApprox-02}(a). Hence, the diagonal $d$ can be mapped to an image curve on or above $d_1^{'}$ within Fr\'{e}chet distance\ $3\varepsilon$.
If this image curve $d_1^{'}$ then crosses another image curve $d_2^{'}$ this argument above cannot be just repeated because the approximation factor would depend linearly on the number of image curves which cross each other. The preimage points on $d$ of such an intersection not involving $d^{'}$ are separated by at most $6\varepsilon$. This is because both diagonals involved in the intersection have a $3\varepsilon$ correspondence between the region of them mapped above $d^{'}$ and $d$. If the preimages are in order there is no problem. If they occur out of order they cause a monotonicity constraint. Fortunately, we can collapse this region on $d$ to the leftmost preimage with $6\varepsilon$ and then map it to the corresponding point on $d_2^{'}$ in $3\varepsilon$ for a total of $9\varepsilon$, see Figure \ref{fig:FoldedApprox-02}(b). If the preimage points in this example were reversed they would be in order.
Thus, we can approximate away the monotonicity constraint of single intersections with $9\varepsilon$. We must also verify that if the preimages of single intersections occur out of order it does not effect our approximation, see Figure \ref{fig:FoldedApprox-02}(c). Due to lack of space the discussion of these technical cases has been moved to the appendix, see Appendix \ref{sec:CFATechnicalCases}.
From these we get that $\delta_F(d,d^{''}) \leq 9\varepsilon$. Now collapse the ear we initially selected in $P$ to $d$. Likewise in $Q$ collapse the corresponding ear to $d^{''}$. This is okay because $d^{''}$ is above all of the other image curves in Q. This pairs the ear we cut off of $P$ with the part of Q above $d^{''}$ which is a folded polygon\ (it could be simpler than a folded polygon\ but we know it's no more complex than that). By Lemma \ref{lem:GeneralApprox-ConvexVSFoldedPolygon} the ear of $P$ and the folded polygon\ above $d^{''}$ must be within Fr\'{e}chet distance\ $9\varepsilon$.
Choose another ear in $P$. We can repeat the above arguments to remove this new ear and its corresponding ear in $Q$. The dual graph of $P$ is a tree. Each time we repeat this argument we are removing a leaf from the tree. Eventually, the tree will contain only a single node which corresponds to some triangle in $P$ which we map to the remainder of $Q$.
\qed
As a direct consequence of Theorems \ref{thm:GeneralApprox-SimpleTest} and \ref{thm:SimpOptimized-Runtime} we get the following theorem:
\begin{theorem}
\label{thm:GeneralApprox-Algorithm}
We can compute a $9$-approximation of the Fr\'{e}chet distance\ of two folded polygons\ in time $O(kT_{matrixmult}(mn)\log(mn))$.
\end{theorem}
\vspace{-3 mm}
\section{Axis-Parallel Folds and $L_{\infty}$ Distance \label{sec:LinfParallel}}
The counter example from Appendix \ref{sec:CounterExample} works for all $L_p$ except for $L_{\infty}$. In this section we outline a special case where using the $L_{\infty}$ metric guarantees that if a pair of folded polygons\ pass the diagonal monotonicity test\ for $\varepsilon$ their Fr\'{e}chet distance\ is no more than $\varepsilon$. Specifically, if all of the line segments in the convex subdivision of the surfaces are parallel to the x-axis, y-axis, or z-axis, we show that it is sufficient to use shortest paths instead of Fr\'{e}chet shortest paths. Since shortest paths never cross we can use the simple polygons algorithm in \cite{Buchin2006} to compute the Fr\'{e}chet distance\ of the surfaces. We first prove the following lemma.
\begin{lemma}
\label{lem:HalfSpaceRestrictions}
Let $R$ be a half-space such that the plane bounding it, $\partial R$, is parallel to the xy-plane, yz-plane, or xz-plane. Given a folded polygon\ $Q$ with edges parallel or perpendicular to the x-axis and points $a,b \in Q \cap R$, let $f$ be a path in $Q$, which follows the shortest path edge sequence between $a$ and $b$. If $P$ is completely inside of $R$ so is the shortest path $f^{'}$ between $a$ and $b$.
\end{lemma}
For the lemma to be false there must exist a $Q$, $R$, and $f$ which serve as a counter example. There must be at least one edge $e_j$ in $Q$ such that $f \cap e_j \in R$ and $f^{'} \cap e_j \not\in R$. In particular, let $e_j$ be the first edge where this occurs along the shortest path edge sequence. First consider a $Q$ where all of the edges of it are perpendicular to $\partial R$. A line segment in the shortest path $f^{'}$ connects the endpoints of two edges in $Q$. Let $e_i$ and $e_k$ be the edges that define the line segment in $f^{'}$ that passes through $e_{j}$. We now consider several cases in how those edges are positioned.
Case (I) occurs when $e_k$ is completely outside of $R$, see Figure \ref{fig:ParallelPerpTangleExpNew-02}(a). While this does force $f^{'}$ to cross $e_j$ outside of $R$, there is no $f$ which can pass through $e_k$ while remaining inside of $R$. Because $Q$ is a folded polygon\ any path between $a$ and $b$ must path through the edges in the shortest path edge sequence including $e_k$. Thus no $f$ can exist entirely within $R$.
Case (II) occurs when part of $e_k$ is in $R$ and $f^{'}$ crosses it in the part in $R$, see Figure \ref{fig:ParallelPerpTangleExpNew-02}(b). In this case $f^{'}$ does not cross $e_j$ outside of $R$.
Due to space limitations the discussion of the remaining cases has been moved to the appendix, see Appendix \ref{sec:Lemma4Appendix}. Each of the remaining cases can be reduced to these first two. Using this lemma we can prove the following theorem:
\begin{theorem}
\label{thm:ParallelPerpCase}
The Fr\'{e}chet distance\ between two surfaces, both with only diagonals/edges parallel to the x-axis, y-axis, or z-axis, can be computed in time $O(kT_{matrixmult}(mn)\log(mn))$.
\end{theorem}
\vspace{-3 mm}
\begin{figure}[ht]
\centering \includegraphics[width=0.90\textwidth]{ParallelPerpTangleExpNew-02}
\centering \caption{\small (a), (b) are examples of case (I) and case (II). (c) example intervals for the two different paths. (d) together the edges $e_1$ and $e_2$ cause a monotonicity constraint.}
\label{fig:ParallelPerpTangleExpNew-02}
\vspace{-4 mm}
\end{figure}
\proof Let $Q$ be a folded polygon, $d$ be a diagonal, and $f^{'}$ be the shortest path between points $a$ and $c$ on $\partial Q$. Using Lemma \ref{lem:HalfSpaceRestrictions} we prove that if there exists a Fr\'{e}chet shortest path\ $f$ between points $a$ and $c$ such that $\delta_F(d,f) \leq \varepsilon$, then $\delta_F(d,f^{'}) \leq \varepsilon$.
\vspace{-3 mm}
\subsubsection*{Minkowski Sum Constraints}
Since we are using the $L_{\infty}$ distance, the unit ball is a cube with sides of length 1. The Minkowski sum of a diagonal $d$ in $P$ and a cube of side length $\varepsilon$ yields a box. Points in the diagonal $d$ can only map to points in this region. It can be defined by the intersection of 6 half-spaces; all of these have boundaries parallel to either the xy-axis, the xz-axis, or the yz-axis. Thus, from Lemma \ref{lem:HalfSpaceRestrictions}, we know that if any path through $Q$ is completely within this box, then the shortest path $f^{'}$ will be, too. This means that for each edge $e_i$ on the shortest path edge sequence $f^{'} \cap e_i$ is within distance $\varepsilon$ of some non-empty interval of $d$.
\vspace{-3 mm}
\subsubsection*{Monotonicity Constraints}
For the shortest path $f^{'}$ between the boundary points to have $\delta_F(d,f^{'}) > \varepsilon$, at least two of these intervals must be disjoint and occur out of order along $d$, see Figure \ref{fig:ParallelPerpTangleExpNew-02}(c). Such a case introduces a monotonicity constraint on $\varepsilon$. If no such intervals existed then we could choose a monotone sequence of points along $d$ such that each point is within distance $\varepsilon$ of an edge and the sequence of edges they map to would have the same order as the shortest path edge sequence showing that $\delta_F(d,f^{'}) \leq \varepsilon$.
Let $e_1$ and $e_2$ be two edges along the shortest path edge sequence for which such bad intervals occur. Let $p_1$ and $p_2$ be points on the shortest path where it intersect edges $e_1$ and $e_2$ respectively. Let $q_1$ and $q_2$ be the same for $f$. Finally, let $d_{r}$ contain all of the points on $d$ which are within distance $\varepsilon$ of the point $r$. Since $\delta_F(d,f) \leq \varepsilon$, $d_{q_1}$ and $d_{q_2}$ must overlap or occur in order along $d$.
Let $R_1$ be the half-space whose bounding plane contains $q_1$ and is perpendicular to $d$. likewise let $R_2$ be the half-space whose bounding plane contains $q_2$ and is perpendicular to $d$, see Figure \ref{fig:ParallelPerpTangleExpNew-02}(d). Let $R_1$ extend to the left along $d$ and $R_2$ extend to the right along $d$. $\partial R_2$ must occur before $\partial R_1$ along $d$ or the edges are in order and no monotonicity constraint is imposed. Assume $R_1$ encloses all of $f^{'}$ between $a$ and $e_2$. If it does not we can choose a new edge between $a$ and $e_2$ to use as $e_1$ for which this is true. Doing so only increases the monotonicity constraint. Likewise we can assume $R_2$ encloses all of $f^{'}$ between $e_2$ and $c$.
Assume $a$, $b$, and $c$ lie on $f^{'}$. Specifically, let $a$ and $c$ be the end points of $f^{'}$ on $\partial Q$. Naturally, a shortest path must exist between $a$ and $c$ and it must contain at least one point in $R_1 \cap R_2$ which we call $b$. $f$ follows the shortest path edge sequence between $a$ and $c$, so it must also cross all of the edges in the shortest path edge sequence between $a$ and $b$. Therefore, to show that $p_2$ is inside of $R_1$ we can directly apply Lemma \ref{lem:HalfSpaceRestrictions} to the points $a$ and $b$. A similar method can be used for $e_2$ with points $b$ and $c$ to show $p_1$ is inside $R_2$. Since $d_{q_1}$ and $d_{q_2}$ overlap or are in order, $d_{p_1}$ and $d_{p_2}$ must as well. Therefore, $\delta_F(d,f^{'}) \leq \varepsilon$ and shortest paths can be used for this variant of folded polygons\ instead of Fr\'{e}chet shortest paths. Because we are using shortest paths we can just use the simple polygons algorithm. This yields Theorem \ref{thm:ParallelPerpCase}. \qed
\vspace{-3 mm}
\section{Future Work\label{sec:Conclusion}}
The constant factor approximation outlined in Section \ref{sec:GeneralApprox} can likely still be improved. Specifically, we consider only the worst case for each of the out-of-order mappings which may not be geometrically possible to realize. In addition, we currently approximate the Fr\'{e}chet distance\ by mapping image curves one-by-one to the top of the arrangement of other image curves. It would of course be more efficient to untangle image curves by mapping them to some middle curve rather than forcing one to map completely above the others.
Finally, while the problem of untangling seems hard, it is also possible that a polynomial-time exact algorithm could exist. The acyclic nature of our surfaces seems to limit the complexity of our mappings. The methods used to prove that computing the Fr\'{e}chet distance\ between certain classes of surfaces is NP-hard in \cite{bbs-fds-10} are not easy to apply to folded polygons.
\vspace{-4 mm}
\bibliographystyle{lipics}\setlength{\itemsep}{-2mm}
|
\section{Introduction}
Ultra-luminous X-ray sources (ULXs) are bright point-like sources found at off-nuclear positions in nearby
galaxies. If the X-ray emission is isotropic, the ULX luminosities would be in the range 10$^{39}$ -- 10$^{41}$ \ergs,
with corresponding Eddington-limited masses in excess of 100 \msol. This leads to the suggestion that ULXs
are black holes with mass intermediate between stellar-mass systems and very massive nuclear black holes of galaxies.
The large luminosities, however, may still originate from stellar-mass black holes if the X-ray emission is beamed,
or if accretion disks are capable of radiating at super-Eddington rates \citep[e.g.,][]{ket01, b02}.
Indeed some X-ray binaries are known to produce super-Eddington luminosities,
reaching \lower.5ex\hbox{$\; \buildrel > \over \sim \;$}\ 10$^{41}$ \ergs\ in their flaring states \citep[e.g.,][]{me01, mew03a, mew03b, ssj05}.
The nature of ULXs is still quite uncertain, and given they are broadly selected as bright off-nuclear
X-ray sources, they may represent a mixture of different types of objects \citep{zr09}.
Many ULXs have early-type stars as optical counterparts, indicating their binary nature.
Some also have extended optical nebulae surrounding the central source \citep[see][for early results]{pm02}.
Of particular interest is the detection of the \heii\ line at 4686 \AA\ (\heii\ $\lambda$4686)
in the immediate vicinity (\lower.5ex\hbox{$\; \buildrel < \over \sim \;$}\ 2\arcsec) of some of bright sources
\citep[e.g.,][]{get06, pm02, kwz04, let05, kc09}.
Because of the very high ionization potential of 54.4 eV,
this line emission traces high-energy (EUV to X-ray) radiation fields or extremely strong shocks,
and is usually considered to be a signpost of X-ray photo-ionized nebulae \citep[e.g.,][]{pa86, pm89}.
Where present, the observed \heiiw\ luminosities have been used to obtain
lower limits to the total X-ray ray luminosity \citep[e.g.,][]{kc09},
which provides a constraint central to determining the nature of ULXs.
In this {\em Letter}, we present deep spectroscopic observations of the \heiiw\ line in three ULXs, \holix; \meo; and \holii.
\holix\ (= M81 X-9) is located in the dwarf galaxy Holmberg IX near M81,
and is known to have variable X-ray emission with a luminosity (assuming isotropic emission)
of $\sim$~10$^{40}$ \ergs\ \citep{lpet01},
along with a large ($\ge$ 250 pc in diameter) optical nebula seen in H recombination and forbidden lines \citep{m95, get06}.
Its optical counterpart, which is probably an early-type star of $V$ $\simeq$ 22.8 \citep{get06},
shows strong, broad, spatially unresolved \heiiw\ line emission \citep{get06}; however \citet{ret06} was unable to detect
extended \heiiw\ emission. \meo\ (= NGC 3031 X-11) resides in the spiral galaxy M81 and
also shows long-timescale variable X-ray emission with an isotropic-equivalent luminosity of $\sim$~2 $\times$ 10$^{39}$ \ergs.
Its optical counterpart is an O8~V star of $V$ $\simeq$ 23.9 situated in a spiral arm with high local stellar density
\citep{rw00, let02} and embedded in an \halpha\ nebula $\sim$~150 pc in diameter \citep{ret06}.
Prior to the observations reported here no \heiiw\ emission has been detected in \meo.
\holii\ is one of the most luminous ($>$ 10$^{40}$ \ergs) ULXs,
and is located in the dwarf star-forming galaxy Holmberg II.
\citet{kwz04} identified a likely optical counterpart that is an early-type star
surrounded by a nearby optical nebula including \heiiw\ emission $\sim$~2\arcsec\ in size.
\citet{let05} confirmed the extended \heiiw\ region of 21 $\times$ 47 pc, and determined
$B$ $\simeq$ 20.05 magnitude of the optical counterpart.
In addition this ULX appears to have a radio counterpart \citep{mmn05}.
The observed \heiiw\ luminosity of 2.7 $\times$ 10$^{36}$ \ergs\ corresponds to a lower limit on the X-ray luminosity
of 4--6 $\times$ 10$^{39}$ \ergs, assuming that X-ray photo-ionization is the source of the line excitation \citep{kwz04}.
\section{Observations and Results}
The observations of the three ULXs, \holix; \meo; and \holii, occurred
2006 January 31, 2006 February 2--3,
and 2007 March 3--5 using the Low Resolution Imaging Spectrograph \citep[LRIS;][]{oet95}
on the Keck~I 10-m telescope.
The $B$-band seeing was 1.5\arcsec -- 2\arcsec\ (2006) and 1.0\arcsec -- 1.5\arcsec\ (2007);
the slit width and length were 1\arcsec\ and 120\arcsec, respectively.
A beam dichroic inside LRIS splits incoming light to the blue and red sides followed by
a 600/4000 grism (3010--5600 \AA) and a 600/7500 grating (5700--8200 \AA).
Photometric calibrations were performed with standard sources HD~93521 and HD~9,
and wavelength solutions were obtained using arc lamp lines.
Table 1 summarizes the observations.
We accumulated multiple exposures of 1800-s each for \holix\ and \meo,
totalling on-source integration time up to 5.5 (\holix) and 7.0 (\meo) hours.
For \holii, which is much brighter,
we integrated for a total of $\sim$ 0.3 hours.
The slit position angles for \holix\ and \holii\ were fixed to be 0$^\circ$,
while two different position angles, 0$^\circ$ and 72.5$^\circ$, were used for \meo.
We detected extended \heii\ emission surrounding all three objects.
The line was not detected in the 0$^\circ$ position angle observations of \meo , however
the signal to noise of these observations makes non-detection consistent with the level seen in the
2006 data. We therefore present the analyses of the 2006 data of \meo\ alone.
We also detected several Balmer series H lines and
forbidden transitions of [\ion{O}{1}], [\ion{O}{2}], [\ion{O}{3}], [\ion{Ne}{3}], [\ion{N}{2}], and [\ion{S}{2}].
Detailed analyses of the entire spectra will be published elsewhere; in this {\it Letter} we concentrate
on the \heiiw\ and \hbeta\ line spatial and spectral analysis.
Figure~\ref{fig_spectrogram} shows integrated spectrograms of the three ULXs
in the wavelength range 4650 -- 4900 \AA, containing the \heii\ ($\lambda$4686)
and \hbeta\ ($\lambda$4861) lines. The offset ($x$-axis) is in arcseconds relative to the reported
position of the optical counterpart in each system. The dispersed continuum emission from the
optical counterparts of \holix\ and \holii\ is clearly visible in Figure~\ref{fig_spectrogram}(a) and (c), respectively.
In the \holii\ spectrogram, in addition to the optical counterpart,
there are two bright background stars between --3\arcsec\ and --7\arcsec.
The counterpart detection in \meo\ in Figure~\ref{fig_spectrogram}(b) is uncertain.
The dispersed light of its optical counterpart, which is fainter relative to the other two ULXs,
appears to be blended with that of many other nearby stars \citep[see][]{let02}.
The $\simeq$ 2\arcsec\ seeing of these observations make the continuum from the counterpart impossible to isolate from
that of nearby stars.
(The bright continuum source at $\sim$ 15\arcsec\ to the right in Figure~\ref{fig_spectrogram}[b] is
a field star used for the slit alignment.)
In all three spectrograms extended \heii\ and \hbeta\ emission is clearly detected.
Figure~\ref{fig_profile} compares the spatial distribution of the integrated
\heii\ and \hbeta\ line intensities (solid line for \heii; dotted line for \hbeta)
along the slit direction, with the same $x$-axis origin used in Figure~\ref{fig_spectrogram}.
The \hbeta\ distribution is shifted by an arbitrary constant (0.3) in the y-direction for clarity.
We subtracted the background, including the stellar continuum, using the flux detected in nearby pixels.
A common feature of all three (especially \holix\ and \holii) profiles is that,
with the exception of the narrow central component in \holix,
the \hbeta\ spatial profile generally tracks,
but is more extended than, the \heii\ emission.
For all the three ULXs, the \heii\ emission extends to $>$100~pc from the center,
while the \hbeta\ emission extends much further, to beyond 250~pc.
The \heiiw\ to \hbeta\ line intensity ratios are
$\sim$0.12 (\holix), 0.19 (\meo), and 0.33 (\holii).
Below we discuss the spatial distribution for each source.
The \heii\ line intensity distribution in \holix\ (Figure~\ref{fig_profile}[a]) can be fit using three gaussian
components; the fit is shown as a dashed line. The central, narrow, unresolved component is described by a gaussian profile with a
FWHM of 1.0\arcsec\ (or 17.5 pc), comparable to the seeing size. Its integrated intensity contains $\sim$1/3 of the total \heii\ emission.
This component is surrounded by asymmetric emission extending more than 100~pc in both directions,
but more prominently to the south (left). The extended \heii\ emission can be fit using two Gaussian components:
one centered at the origin with FWHM of 8.7\arcsec\ ($\simeq$ 150 pc),
the other at --4.3\arcsec\ ($\simeq$ 75 pc) with FWHM of 2.4\arcsec\ ($\simeq$ 40 pc).
The first component is roughly 3 times brighter than the second one,
and the intensity of the total emission in the south is greater than that in the north by $\sim$40\%.
We estimate the extent of the \heii\ emission to be $184 \pm 20$~pc, determined by
measuring the continuous area across the center where the signal intensity of each
data point is greater than noise level at the 90~\% confidence level.
We estimated the noise level at $<$ --15\arcsec\ and $>$ +15\arcsec,
where there is no apparent \heii\ emission.
The \hbeta\ emission of this source is asymmetric without any apparent central peak.
With the notable exception
of the central peak, however, the \hbeta\ profile is similar in shape to that of the \heii.
The \hbeta\ spreads somewhat flatly to $\sim$300~pc in the northern direction,
whereas, in the south, it extends less ($\sim$200~pc) and peaks
where the \heii\ emission has a local peak. The total extension of the \hbeta\ emission, therefore, is $\sim$500 pc,
and it is slightly ($\sim$10~\% ) brighter in the south than the north.
We estimate the total luminosity of the \heii\ line emission
integrated along the slit to be $\sim$ 3.7 $\times$ 10$^{35}$ \ergs\ for a distance of 3.6 Mpc.
There is no identifiable central component to the \heii\ emission in \meo\ (Figure~\ref{fig_profile}[b]).
The distribution is irregular, and not well described by gaussian components.
The \heii\ is extended $115 \pm13$~pc west-southwest (left) from the center;
whereas it extends $42 \pm 7$~pc in the opposite direction, for a total width of $167 \pm 15$~pc.
Both the \heii\ and \hbeta\ emission is heavily ($>$ 90~\%) concentrated in the west-southwest
where the \hbeta\ emission extends more than 150~pc.
The \heii\ and \hbeta\ emission is again spatially correlated; however,
the peak locations of the \heii\ emission appear to be shifted toward the center compared to \hbeta.
The main peak of the \heii\ emission may be dividable into two components at $x \simeq 4.5$\arcsec,
although the low signal-to-noise ratio makes it difficult to confirm this.
The total measured \heii\ luminosity is $\sim 8.8 \times 10^{34}$~\ergs\ for a distance of 3.6 Mpc.
For \holii\ (Figure~\ref{fig_profile}[c]), the \heii\ emission is well-described
by a broad, central Gaussian component of 2.7\arcsec\ ($\simeq$ 40 pc) FWHM,
along with a minor component of 1.9\arcsec\ ($\simeq$ 30 pc) FWHM
located at --3\farcs5 ($\simeq$ 50 pc) to the south.
The integrated intensity of the former is roughly 10 times greater than that of the latter.
The \hbeta\ emission is distributed very similarly to the \heii\ emission:
the main peak is at the center and there is a secondary peak near the location
of the secondary component of the \heii\ emission at --3\farcs5.
In addition, there is a third peak centered at --10\arcsec\ in the \hbeta\ emission.
The \heii\ emission is somewhat enhanced at this location,
although it is almost indistinguishable from the noise.
The size of the \heii\ emission is 122 $\pm$ 7 pc.
The measured \heii\ luminosity is $\sim$ 3.6 $\times$ 10$^{36}$ \ergs,
which is slightly greater than that of previous measurements \citep{pm02, kwz04, let05},
for a distance of 3.1 Mpc.
Figure~\ref{fig_line} shows spectral line profiles of the \heii\ (left panels)
and \hbeta\ (right panels) transitions for the three ULXs. There are two \heii\ line profiles for \holix\ and \holii:
the thick-solid profiles are for the central emission obtained within 1\arcsec\ from the center
(i.e., the optical counterparts); the thin-solid ones are for the extended emission outside the center.
The FWHM of the \heii\ lines of the spatially extended emission are
5.0 (\holix), 3.7 (\meo), and 3.5 (\holii) \AA.
Those for the central emission are 6.7 \AA\ (\holix) and 3.6 \AA\ (\holii).
For the \hbeta\ lines, the line widths of the extended emission are
4.4 \AA\ (\holix), 3.6 \AA\ (\meo), and 3.5 \AA\ (\holii).
The LRIS instrumental line widths, estimated from
the measured widths of the calibration lamp lines,
are in the range 3.8 -- 4.1 \AA.
We therefore conclude that the \heii\ emission of \holix, especially the central component,
has a contribution from dynamical motion in the source,
whereas we can only determine an upper limit of $\sim$ 250 \kms\ for the velocity dispersion of the other lines.
For \holix, the extra widths correspond to velocity dispersions of
370 \kms\ and 230 \kms\ for the central and extended components, respectively.
These dispersions of the \heii\ lines are consistent with previous observations \citep{get06},
and are larger than those reported in \oiiiw\ and \siiw\ lines \citep{am08}.
For \holii, the upper limit is consistent with the results of previous observations \citep{let05}.
\section{Discussion and Conclusions}
Our deep Keck observations spectroscopically identified
spatially extended, highly-ionized \heii\ emission around three ULXs.
In the case of \holix\ and \meo\ this is the first reported
detection of diffuse \heii\ emission from these systems, and in the case of \holii, we find the highly-ionized region to be larger than
previously reported. The sizes we find are in the range of 100 -- 200 pc (in diameter), larger
than any previously-known \heii\ emission around a compact source. The \heii\ nebulae around extremely hot stellar sources,
such as planetary nebulae or stellar X-ray sources,
are generally smaller than 10 pc.
Previous optical observations of some ULXs, including \holix\ and \holii\ observed in this study, detected \heii\ emission only
from the locations of the optical counterparts \citep[e.g., M101 X-1, NGC 1313 X-1, and \holix;][]{ket05, pet05, get06}
or from their close ($<$ 50 pc) vicinities \citep[e.g., \holii\ and NGC 5408 X-1;][]{get06, kc09}.
The identification of large \heii\ emission regions surrounding all three sources included in this study indicates
that extended, highly ionized nebulae are a common or even ubiquitous feature of ULXs. Previous non-detections
likely result from the limited depth of the observations; with Keck LRIS being significantly more sensitive than
other telescopes for this purpose. In all three cases the
extended emission contains the majority of the total \heii\ flux. In the case of \holix\ the central unresolved component
amounts to only 1/3 of the total emission, and there is no apparent central component at all in \meo.
In \holii\ the \heii\ emission is distributed in a relatively broad ($\sim$ 40 pc FWHM), but very regularly distributed
component with a minor contribution from very extended emission in the south.
A distinct feature of the \heii\ emission in \holix\ is the unresolved central component of a significant
($\simeq$ 370 \kms) velocity dispersion without any apparent \hbeta\ counterpart.
One possible explanation is that the central component represents a photo-ionized accretion disc rotating
around the central X-ray source \citep{get06}.
The line broadening in this case is due to rotational motion,
and the lack of corresponding \hbeta\ emission suggests that it is gas of relatively small column density.
The large velocity dispersion implies that the disk size
is much smaller than the seeing size for any conceivable mass range for the central source, consistent with our results (\S~2).
As noted above, \meo\ lacks a central component altogether.
In \holii\ the absence of a large velocity dispersion in the central region
(offset of $\leq$ 1\arcsec) probably indicates the lack of an accretion disc,
although it may be due to a projection effect, i.e., \holii\ is close to a face-on system.
The more likely scenario is that the extended
highly ionized nebula is regularly distributed around the system's center and smoothly connects to the broad component.
Another possibility of the unresolved central \heii\ component of \holix\ is that it is due
to strong \heii\ emission from the optical (stellar) counterpart of the source.
For instance, some Wolf-Rayet stars are known to produce luminous \heii\ emission of significant
velocity dispersion \citep[e.g.,][]{ch06}.
We need more information on the optical counterpart to investigate this scenario further.
Our observations point to photo-ionization by intense X-ray flux being the source of the \heii\ and \hbeta\ emission.
The alternative explanation is that both are produced by the same radiative shocks.
However, in the case of \meo\ and \holii\ the observed line intensity ratios of \heii\ to \hbeta\
require much greater velocities than the measured upper limits \citep{aet08},
making it incompatible with radiative shocks.
For \holix, the observed line intensity ratio requires comparable,
but still slightly greater, velocities than observed ($\sim$230 \kms\ for \heii) if the emission is produced in strong shocks.
Therefore, our results favor the interpretation that the \heii\ line emission is dominated by X-ray photo-ionization.
Under this scenario, the smaller extent of the \heii\ emission traces the locations where most of the energetic X-ray
radiation from the ULXs is locally absorbed.
Under the assumption of photo-ionization, the observed \heii\ luminosities can be used to obtain independent measurements
of the ULX X-ray luminosities. The X-ray luminosities estimated based on the observed \heii\ line luminosities
are independent of the assumption that the X-ray radiation from the central source is isotropic
-- the inevitable assumption to calculate the X-ray luminosities from the observed X-ray fluxes.
Measurements to-date based on the line emission all indicate luminosities
in excess of 10$^{39}$ \ergs. For example, the observed \heii\ line luminosity of $\sim$1 $\times$ 10$^{36}$ \ergs\
of the ULX in NGC 5408 provides a lower limit of $\sim$2.5 $\times$ 10$^{39}$ \ergs\
for its X-ray luminosity \citep{kc09} and
the \heii\ luminosity of $\sim$2.7 $\times$ 10$^{36}$ \ergs\ leads to an estimated X-ray luminosity
of $\sim$5 $\times$ 10$^{39}$ \ergs\ in \holii.
In our observations we use a relatively small (1\arcsec\ in width) slit, so that the measured luminosities given in
\S~2 are firm lower limits.
One way to estimate the upper limits on the extended \heii\ line luminosities is
to scale up the observed luminosities within the slit under the hypothesis that
the \heii\ emission is uniformly and symmetrically distributed outside the slit.
In doing so we obtain the upper limits of $\sim$1.2 $\times$ 10$^{37}$ \ergs\ and
$\sim$4.4 $\times$ 10$^{36}$ \ergs\ for \holix\ and \meo, respectively,
within an 8\arcsec\ radius.
We then calculate the X-ray luminosities required to produce the \heii\ line luminosities
using the photo-ionization modelling program $CLOUDY$ \citep[][ver. 07.02.01]{fet98}.
In the modelling, we used an input X-ray spectrum of a multicolor disc blackbody
plus a cutoff power law component
for \holix\ \citep{dgr06}, and disk blackbody component for \meo\ \citep{set03}.
We also assumed a constant gas density of 10 cm$^{-3}$, a unity filling factor,
and a spherically symmetric geometry.
We find that to account for the \heii\ line luminosities the required X-ray luminosities are
in the range of $\sim$10$^{39}$ -- 10$^{40}$ \ergs,
comparable to the observed ULX X-ray luminosities.
This favors models where the X-ray emission is isotropic rather than beamed,
which includes the scenario where ULXs are intermediate mass black holes.
\acknowledgments
D.-S.M. acknowledges the support by NSERC through Discovery program 327277 and
S.B.C. acknowledges generous support from Gary and Cynthia Bengier and the Richard and Rhoda Goldman fund.
The data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among
the California Institute of Technology, the University of California, and National Aeronautics and Space Administration. The
Observatory was made possible by the generous financial support of the W.M. Keck Foundation.
{\em Facilities:} Keck:I (LRIS)
|
\section{The three-site Higgsless model}
\label{sec:3site-review}
In this section we review
the three-site Higgsless model
briefly.
This is a minimal Higgsless model, and its electroweak gauge symmetry is
$SU(2)_0 \times SU(2)_1 \times U(1)_2$. The electroweak symmetry breaking is
described using the Hidden local symmetry language, or non-linear sigma
fields.
We explain gauge and fermion sectors with its Lagrangian
and physical features for each sector.
\input{./3sHLM/gauge_sec.tex}
\input{./3sHLM/fermion_sec.tex}
\input{./3sHLM/couplings_general.tex}
\subsection{Couplings among the heavy gauge boson and light fermions}
\label{sec:constraint_at_tree}
Coupling among $W'$ and light fermions, namely $g_{W'ff}$,
is strongly constrained from electroweak precision measurements \cite{Abe:2008hb}.
Its order of magnitude is
$g_{W' ff}/g_W \sim {\cal O}(10^{-2})$,
where $g_W = e/s_Z $
and $s_Z^2 \equiv 1 - c_Z^2$,
$c_Z \equiv M_W/M_Z$.
Coupling among $Z'$ and light fermions are as follows.
\begin{align*}
g_{Z' f_L f_L}
&\simeq
g_{W}
\left(
\frac{\tau^3}{2} G_3
-
t_Z^2 \frac{M_{W}}{M_{W'}}
Q
\right)
,
\\
g_{Z' f_R f_R}
&\simeq
g_{W}
\left(
-
t_Z^2 \frac{M_{W}}{M_{W'}}
Q
\right)
,
\end{align*}
where
\begin{align*}
G_3
&\equiv
\sqrt{2}
\frac{g_{W' ff}}{g_W}
+
t_Z^2 \frac{M_{W}}{M_{W'}}
,
\\
t_Z
&\equiv
\frac{s_Z}{c_Z}
.
\end{align*}
We can see that
$g_{Z' ff}$ is larger than $g_{W'ff}$,
and
its parameter dependence is moderate compared with $g_{W'ff}$.
Therefore $Z'$ is more suitable than $W'$ as a discovery channel
via DY production process.
\subsection{Fermion sector}
This model has following fermions;
$\Psi _{L0} \sim (2,1)_{Y}$,
$\Psi _{L1} \sim (1,2)_{Y}$,
$\Psi _{R1} \sim (1,2)_{Y}$
and
$\Psi _{R2} \sim (1,1)_{Q}$,
where
$Y = 1/6$ for quarks
and
$Y = -1/2$ for leptons.
We can write down kinetic terms of fermion by using them.
The representation of fermions in this model is summarized in Table
\ref{table:fermion_rep}.
\begin{table}[H]
\begin{center}
\begin{tabular}{c|cccc}\hline\hline
& $SU(2)_0$ & $SU(2)_1$ & $U(1)_2$ & $SU(3)_c$\\ \hline
$\Psi_{L0}$ & $2$ & $1$ & $\frac{1}{6}$
$\left(-\frac{1}{2}\right)$ & $3$ ($1$) \\
$\Psi_{L1}$ & $1$ & $2$ & $\frac{1}{6}$
$\left(-\frac{1}{2}\right)$ & $3$ ($1$)\\
$\Psi_{R1}$ & $1$ & $2$ & $\frac{1}{6}$
$\left(-\frac{1}{2}\right)$ & $3$ ($1$)\\
$\Psi_{R2}
=\left(
\begin{array}{ccc}
u_{R2} \\
d_{R2} \\
\end{array}
\right) $ & $1$ & $1$ & $
\begin{array}{ccc}
\frac{2}{3} \\
-\frac{1}{3} \\
\end{array}
$ $\left(
\begin{array}{ccc}
0 \\
-1 \\
\end{array}
\right) $ & $3$ ($1$)\\
\hline\hline
\end{tabular}
\caption{The representation of fermions in this model.
2 and 3 mean fundamental representations for $SU(2)$ and $SU(3)$
respectively, and 1 means singlet.
Other numbers are values of hypercharge.
The numbers in
parenthesis in columns of $U(1)_2$ and $SU(3)_c$ are for leptons.}
\label{table:fermion_rep}
\end{center}
\end{table}
Mass and Yukawa interaction terms in this model are as follows;
\begin{eqnarray*}
{\cal L}_{\mathrm{fermion}}^{\mathrm{Yukawa}}
=
-
\sum_{i,j}
&\Biggl[&
(\overline{\Psi}_{L0})^{i} U _1 m_{1 ij} (\Psi_{R1})^{j}
\\
&&
+ (\overline{\Psi}_{L1})^{i} M_{ij} (\Psi_{R1})^{j}
\\
&&
+ (\overline{\Psi}_{L1})^{i} U_2 m_{2 ij} (\Psi_{R2})^{j}
\\
&&
+ (h.c) \frac{}{}
\Biggr],
\label{eq:lagrangian-Yukawa-sec}
\end{eqnarray*}
where
$M$ is Dirac mass,
$i$ and $j$ are indices of generation or flavor,
and
\begin{equation*}
m_{2 ij}
=
\left(
\begin{array}{cc}
m_{2u} & 0 \\
0 & m_{2d} \\
\end{array}
\right)_{ij}.
\end{equation*}
In general, $m_{1ij}$ and $M_{ij}$ are not flavor blind.
However, to avoid a large FCNC,
$m_{1ij}$ and $M_{ij}$ are assumed to be flavor blind, namely,
$m_{1ij} = m \delta_{ij}$ and $M_{ij} = M \delta_{ij}$.
Under this assumption, the structure of all flavors in this model is embedded in
$m_{2u}$ and $m_{2d}$.
Again, $\Psi$'s are not mass eigenstates but gauge eigenstates.
By diagonalizing mass matrices,
we find that masses of heavy fermions can be described to be approximately $M$.
Using constraints from $S$ and $T$ parameters,
we can find a lower bound on $M$,
$M \geq 1800$ GeV, which is much heavier than bounds for heavy gauge bosons.
Therefore the study of production processes of heavy gauge bosons,
is more promissing than heavy fermions at the LHC.
\subsection{Gauge sector}
\label{3site-gauge-sec}
The gauge sector of this model is written as
\begin{eqnarray}
{\cal L}_{\mathrm{gauge}}
&=&
-\frac{1}{4} G_{\mu\nu}^a G^{a\mu\nu}
\label{eq:3site-gluon}
\\
&&
-\frac{1}{4}\sum_{i=0,1} W_{i\mu\nu}^a W_i^{a\mu\nu}
-\frac{1}{4}B_{\mu\nu} B^{\mu\nu}
\nonumber
\\
&&
+\sum_{i=1,2}\frac{f_i^2}{4}\mbox{tr}\bigl[
(D_\mu U_i)^\dagger (D^\mu U_i)
\bigr] .
\label{eq:3site-gauge-sec}
\end{eqnarray}
Eq.~(\ref{eq:3site-gluon}) is a gluon sector,
and Eq.~(\ref{eq:3site-gauge-sec}) are a electroweak sector.
$W_{0 \mu}$, $W_{1 \mu}$ and $B_{\mu}$ are gauge fields of $SU(2)_0$,
$SU(2)_1$ and $U(1)_2$, respectively. Their gauge couplings are $g_0$,
$g_1$ and $g_2$, respectively.
$U_{i}$ are would-be Nambu-Goldstone bosons in non-linear sigma
representation\footnote{$\tau^a$ is Pauli matrices.},
\begin{eqnarray*}
U_i &=& \exp \left(i\frac{\pi^a_i \tau^a}{f_i}\right),
\end{eqnarray*}
and their covariant derivatives are following;
\begin{eqnarray*}
D_\mu U_1 &=& \partial_\mu U_1
+ i g_0 \dfrac{\tau^a}{2} W^a_{0\mu} U_1
- i g_1 U_1 \dfrac{\tau^a}{2} W^a_{1\mu} \\
D_\mu U_2 &=& \partial_\mu U_2
+ i g_1 \dfrac{\tau^a}{2} W^a_{1\mu} U_2
- i g_2 U_2 \dfrac{\tau^3}{2} B_{\mu} .
\end{eqnarray*}
Notice that $U(1)_2$ gauge symmetry acts on $U_2$ as a ``partially gauged
$SU(2)_2$''.
All fields in the above Lagrangian are not mass eigenstate but gauge
eigenstate except gluon field.
Mass matrices for gauge bosons can be derived from Eq.~(\ref{eq:3site-gauge-sec}).
By taking a mass eigenstate basis, we can obtain physical particles, that is,
charged gauge bosons~($W^{\pm}_{\mu}$ and $W^{\prime \pm}_{\mu}$)
and neutral gauge bosons~($Z_{\mu}$, $Z'_{\mu}$ and $A_{\mu}$)
as linear combinations of $W_{0 \mu}$, $W_{1 \mu}$ and $B_{\mu}$.
There are three extra heavy gauge bosons $W^{\prime \pm}$ and $Z'$,
which have almost the same mass.
From experimental constrains on
the $WWZ$ coupling \cite{Hagiwara:1986vm, :2005ema},
we get a lower bound on them,
$M_{Z'}\sim M_{W'} \geq 380$ GeV.
Using constraints from $S$ and $T$ parameters,
we can find a upper bound on them,
$M_{Z'}\sim M_{W'} \leq 610$ GeV.
\section*{Acknowledgments}
T.A. is supported in part by the JSPS Grant-in-Aid No. 204354.
\section{Discovery Potential at the LHC}\label{sec:analysis}
In this section, we perform feasibility studies of the Higgsless model at the LHC.
To investigate the $Z'$ discovery potential,
we apply a simple detector simulation with smearing methods,
which approximately reproduce the ATLAS experimental condition at proton-proton collision
of a center-of-mass energy of $\sqrt{s}=14$~TeV~\cite{Aad:2009wy}.
The discovery potential is studied with $Z'\rightarrow WW\rightarrow \ell\nu qq$~($\ell=e$, $\mu$) decay process,
where one of $W$ bosons decays leptonically and the other hadronically.
This channel is capable of reconstructing $Z'$ resonance by solving analytically longitudinal component of a neutrino as described in the following section.
Finally, the discovery potential estimated from invariant mass of two $W$ bosons, $M_{WW}$, as a function of integrated luminosity is shown.
\subsection{MC sample and cross section}
Signal and dominant background processes are generated with various Monte Carlo~(MC) generators as follows.
This study uses three preferable $Z'$ signal mass points whose parameters are summarized in Table~\ref{table:sample-points}.
The $Z'$ signal and $WW$ background processes are generated with CalcHEP~\cite{calchep} and
the parton shower and hadronization are simulated with PYTHIA~\cite{pythia}.
The $t\bar{t}$ process is generated with MC@NLO~\cite{mcatnlo}, $W$+jets with ALPGEN~\cite{alpgen,MLM}
and the parton shower and hadronization are simulated with HERWIG~\cite{herwig}.
The cross section of $Z'$ signal and background processes is summarized in Table~\ref{tbl:Xsec}.
The ATLAS detector effects are taken into account by
the Monte Carlo simulation smeared with a simplified ATLAS detector~\cite{AcerDET}.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{ccccc}
\hline \hline
$M_{Z'}$(GeV) & $M$ (GeV) & $\frac{g_{W'ff}}{g_W}$ & Br($Z' \to WW$) \\ \hline
380 & 3500 & 0.023 & 0.971 \\ \hline
500 & 3500 & 0.023 & 0.987 \\ \hline
600 & 4300 & 0.022 & 0.991 \\ \hline\hline
\end{tabular}
\end{center}
\caption{Parameters of $Z'$ signal samples}
\label{table:sample-points}
\end{table}
\begin{table}[htbp]
\begin{center}
\begin{tabular}{cc} \hline\hline
\multicolumn{2}{c}{Signal Sample}\\
$M_{Z'}$ (GeV) & Cross section (pb) \\ \hline
380 & 5.886\\
500 & 4.150\\
600 & 4.165\\ \hline
\multicolumn{2}{c}{Background Sample}\\
Process & Cross section (pb) \\ \hline
$WW$ & 111.6~\cite{Aad:2009wy,MCFM} \\
$t\bar{t}$ & 883~\cite{Langenfeld:2009wd} \\
$W(\rightarrow \ell\nu)$+jets & 20510~\cite{Aad:2009wy,PhysRevLett.96.231803} \\
\hline \hline
\end{tabular}
\caption{Cross section of $Z'$ signal and the Standard Model background processes.
The number of $W$+jets is for single lepton flavor.}
\label{tbl:Xsec}
\end{center}
\end{table}
\subsection{Event selection}
The final state of $Z'$ considered in this paper is $\ell\nu qq$.
Experimentally, the neutrino can be observed as a missing transverse energy~($E_{\mathrm{T}}^{\mathrm{miss}}$) and
the quark is observed as a jet which is a cluster of hadron.
While the lepton ($e$ and $\mu$) can be measured precisely and used for an event trigger.
First, exactly one high-$p_{\mathrm{T}}$\ lepton into the detector coverage of a tracking detector ($p_{\mathrm{T}}^{\ell} > 50$ GeV, $|\eta| < 2.5$)\footnote{A pseudo-rapidity $\eta$ is defined by $-2\ln(\tan{\frac{\theta}{2}})$, where $\theta$ is the angle with respect to the beam line.} is required.
The inefficiency of lepton identification and trigger is taken into account and
80\% of efficiency from combined identification and trigger, which is based on MC studies for the ATLAS detector~\cite{Aad:2009wy}, is applied.
Next, a large missing transverse energy ($E_{\mathrm{T}}^{\mathrm{miss}}$$\ > 50$~GeV) is required.
In addition, exactly two jets with $p_{\mathrm{T}}$$\ > 50$~GeV and $|\eta| < 3.2$ which is corresponding to the coverage of the calorimeter are required.
If there are jets with $p_{\mathrm{T}}$$\ > 25$ GeV, $|\eta| < 2.5$ and matched to $b$-quark,
the $b$-tagging which has 50\% efficiency and 2.5$\times 10^{-3}$ false tag rate are applied.
Here we assume that there is no dependence of $p_{\mathrm{T}}$\ and $\eta$ on the $b$-tagging efficiency and false tag rate.
Events are rejected to reduce enormous $t\bar{t}$ background if at least one $b$-tagged jet exists in the event.
The reconstructed dijet invariant mass, $M_{jj}$, is required to be close to the nominal $W$ mass;
\[ |M_{jj}-M_{W}| < 15~{\rm GeV}.\]
This selection is very effective to suppress $W(\rightarrow \ell\nu)$+jets background which does not have hadronic decay of $W$ boson.
Two $W$ bosons decayed from heavy $Z'$ boson are highly boosted.
Hence the events with high $p_{\mathrm{T}}^{\ell\nu}$~(from $p_{\mathrm{T}}^{\ell}$ and $E_{\mathrm{T}}^{\mathrm{miss}}$) and $p_{\mathrm{T}}^{jj}$ are selected.
The selection criteria depending on $Z'$ mass~($M_{Z'}$) are applied to maximize the discovery potential;
$p_{\mathrm{T}}^{\ell\nu} > 150$, 200 and 250 GeV and $p_{\mathrm{T}}^{jj} > 150$, 200 and 250 GeV are required for $M_{Z'}$ = 380, 500 and 600 GeV, respectively.
$Z'$ invariant mass cannot be reconstructed from observables due to the missing information of
longitudinal neutrino momentum ($p_{z}^{\nu}$).
However we can calculate the longitudinal neutrino momentum by assuming the on-shell $W$ mass constraint\cite{Bach:2009zz};
\begin{equation}\label{eq:WmassConst}
M_{W}^{2} = (E^{\ell}+E^{\nu})^{2}-({\bm p}^{\ell}+{\bm p}^{\nu})^{2},
\end{equation}
where $E^{\ell,\nu}$ and ${\bm p}^{\ell,\nu}$ are energy and momentum of the charged lepton and neutrino, respectively.
The longitudinal component of neutrino momentum is solved analytically from Eq.~(\ref{eq:WmassConst}) as;
\begin{equation}\label{eq:pz}
p_{z}^{\nu} = \frac{(M_{W}^{2}-M_{\ell}^{2}+2{\bm p}_{\mathrm{T}}^{\ell}\cdot{\bm p}_{\mathrm{T}}^{\rm miss})p_{z}^{\ell} \pm \sqrt{D}}{2(E^{\ell 2}-p_{z}^{\ell 2})}
\end{equation}
where ${\bm p}_{\mathrm{T}}^{\ell}$ and ${\bm p}_{\mathrm{T}}^{\rm miss}$ are transverse momentum of charged lepton and neutrino and $D$ is a discriminant represented as
\begin{eqnarray*}\label{eq:D}
D= E^{\ell 2}\left\{(M_{W}^{2}-M_{\ell}^{2}+2{\bm p}_{\mathrm{T}}^{\ell}\cdot{\bm p}_{\mathrm{T}}^{\rm miss})^{2} \right. \\
\left. -4(E_{\mathrm{T}}^{\rm miss})^{2}(E^{\ell 2}-p_{z}^{\ell 2})\right\}
\end{eqnarray*}
Two solutions of $p_{z}^{\nu}$ can be obtained from Eq.~(\ref{eq:pz}).
It is found that a lower $|p_{z}^{\nu}|$ solution have slightly higher probability to match to the true $p_{z}^{\nu}$ value and
gets better $M_{Z'}$ resolution as shown in Table~\ref{tbl:Solution}.
However a higher $|p_{z}^{\nu}|$ solution also has sufficiently high probability to match to the true $p_{z}^{\nu}$.
In addition, 25\% of events have no solution due to a negative discriminant due to the resolution effect of the smearing.
In this case $D$ is likely close to zero. It is found that $p_{z}^{\nu}$ value corresponding to true value
can be obtained even if the imaginary part is neglected in the $p_{z}^{\nu}$ calculation~($D=0$).
Figure~\ref{fig:Solution}~(left) shows the $\Delta M_{WW}$ distribution for each neutrino solution type in $M_{Z'}$ = 500 GeV,
where $\Delta M_{WW}$ is a difference between a reconstructed $M_{WW}$ and its true value.
Since we adopt all the solutions in any case,
the correctly reconstructed one has a peak around zero in the $\Delta M_{WW}$ distribution
but the wrongly one makes a tail in the high $\Delta M_{WW}$ region.
This behavior is observed in the reconstructed $M_{Z'}$ distribution as shown in Figure~\ref{fig:Solution}~(right).
We see not only a clear peak but also a long tail in the high mass region.
Table~\ref{tbl:Solution} shows mean and $\sigma$ values for each solution type.
The $M_{WW}$ distribution of a lower $|p_{z}|$ gives the best resolution and
$M_{WW}$ distribution of other solution types can be reconstructed with slightly higher mean value.
In this study, all three solutions are used to maximize a signal acceptance.
\begin{table}[htp]
\begin{center}
\begin{tabular}{crccc}\hline \hline
Solution Type & \multicolumn{2}{c}{Fraction} & Mean (GeV) & $\sigma$ (GeV) \\ \hline
\multirow{2}{*}{Two solutions} &\multirow{2}{*}{75\% $\Bigl\{$}&55\% (lower $|p_{z}|$)& 500.0 & 18.1 \\
& &45\% (higher $|p_{z}|$)& 503.2 & 21.8 \\
No solution & 25\%& & 507.2 & 19.6 \\ \hline \hline
\end{tabular}
\caption{The fraction of each solution type to events after event selection and mean and $\sigma$ values extracted from a single Gaussian fit for each solution type.}
\label{tbl:Solution}
\end{center}
\end{table}
\begin{figure}[htp]
\begin{center}
\includegraphics[width=4cm,height=4cm]{figures/DelMww_resol_L200M200.eps}
\includegraphics[width=4cm,height=4cm]{figures/Mww_resol_L200M200.eps}
\caption{Left: $\Delta M_{WW}$(reconstructed-true) for each solution type of longitudinal neutrino momentum. Right: Reconstructed $M_{WW}$ distribution for $M_{Z'}$ = 500 GeV.}
\label{fig:Solution}
\end{center}
\end{figure}
\subsection{Discovery potential}
The discovery potential of $Z'$ signal in the three-site Higgsless model is evaluated for three representative mass points.
The number of signal events are defined as
\begin{equation*}\label{eq:Nsig}
N_{\rm signal} = N_{{\rm low}\, |p_{z}|}^{2\, {\rm sol}}+N_{{\rm high}\, |p_{z}|}^{2\, {\rm sol}} + N_{\rm real\, part}^{ {\rm No\, sol}},
\end{equation*}
where $N_{{\rm low}\, |p_{z}|}^{2\, {\rm sol}}$ is for a lower $|p_{z}|$ solution,
$N_{{\rm high}\, |p_{z}|}^{2\, {\rm sol}}$ for a higher $|p_{z}|$ and $N_{\rm real\, part}^{ {\rm No\, sol}}$ for the no-solution case.
Similarly, the number of background events~($N_{\rm bkg}$) are defined as the sum of three solutions.
Figure~\ref{fig:Mww} shows the reconstructed $M_{WW}$ distribution before $p_{\mathrm{T}}^{\ell\nu}$ and $p_{\mathrm{T}}^{jj}$ selection. All neutrino solutions are filled in this plot.
Figure~\ref{fig:Mww500} shows the reconstructed $M_{WW}$ distribution after $p_{\mathrm{T}}^{\ell\nu}$ and $p_{\mathrm{T}}^{jj}$ selection in $M_{Z'}$ = 500 GeV.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=7cm,height=6cm]{figures/Mww_Final.eps}
\caption{Reconstructed $WW$ invariant mass distribution before applying $p_{\mathrm{T}}^{\ell\nu}$, $p_{\mathrm{T}}^{jj}$ selection. All three neutrino solution types are filled in this plot.}
\label{fig:Mww}
\end{center}
\end{figure}
\begin{figure}[htp]
\begin{center}
\includegraphics[width=7cm,height=6cm]{figures/Mww_all_L200M200.eps}
\caption{Reconstructed $WW$ invariant mass distribution after applying $p_{\mathrm{T}}^{\ell\nu}$, $p_{\mathrm{T}}^{jj}$ selection in $M_{Z'} = 500$ GeV. All three neutrino solution types are filled in this plot.}
\label{fig:Mww500}
\end{center}
\end{figure}
The discovery potential is evaluated from the number of expected signal and background in a $M_{Z'}$ mass window, which is determined to get the maximum significance.
This significance is defined as follows;
\begin{equation*}\label{eq:Significance}
{\rm Significance} = \frac{N_{\rm signal}}{\sqrt{N_{\rm bkg}}}\;.
\end{equation*}
Table~\ref{tbl:FinalNumber} shows the number of expected signal and background into the signal mass window for each signal mass point.
The bottom line shows significance. The number of events are normalized to 1~fb$^{-1}$.
\begin{table}[htp]
\begin{center}
\begin{tabular}{cccc}\hline \hline
\multicolumn{4}{c}{Expected cross section with 1~fb$^{-1}$ ($\sqrt{s} = 14$~TeV)} \\ \hline
$M_{Z'}$ (GeV) & 380 & 500 & 600 \\\hline
Optimized Mass window (GeV) & 360-400 & 460-540 & 580-660 \\ \hline
$W$+jets & 53.7 & 45.9 & 9.25 \\
$t\bar{t}$ & 19.8 & 19.9 & 4.75 \\
$WW$ & 5.73 & 5.91 & 1.48 \\
Total background & 79.2 & 71.7 & 15.5 \\ \hline
Signal & 25.5 & 15.3 & 4.6 \\ \hline
Significance & 2.86 & 1.81 & 1.17\\ \hline \hline
\end{tabular}
\caption{The number of signal and background in the signal mass window region for each mass point and obtained significance at 1~fb$^{-1}$. These numbers include all three solutions.}
\label{tbl:FinalNumber}
\end{center}
\end{table}
Figure~\ref{fig:FinalResult} shows the expected significance as a function of integrated luminosity for each mass point.
The significance reaches 3$\sigma$ threshold for $M_{Z'}$ = 380, 500 and 600 GeV in about 1~fb$^{-1}$, 2.5~fb$^{-1}$\ and 6~fb$^{-1}$, and 5$\sigma$ threshold for $M_{Z'}$ = 380, 500 and 600 GeV in about 3~fb$^{-1}$, 8~fb$^{-1}$\ and 20~fb$^{-1}$, respectively.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=7cm,height=6cm]{figures/FinalSensitivity.eps}
\caption{Discovery potential for each $M_{Z'}$ sample as a function of integrated luminosity. Two vertical lines correspond to the 3$\sigma$ evidence and 5$\sigma$ discovery thresholds, respectively.}
\label{fig:FinalResult}
\end{center}
\end{figure}
\section{Introduction}
The standard model (SM) describes the phenomenology of elementary
particles very well.
Its predictions are consistent with many experimental results.
Spontaneous symmetry breaking (SSB) is an important concept in the SM,
and
the electroweak symmetry breaking (EWSB) is known to be spontaneously
broken.
In the SM, the EWSB is triggered by a Higgs boson.
However,
the Higgs boson has not been discovered yet in any experiment.
This fact means the origin of EWSB still remains a mystery.
A Large Hadron Collider experiment at CERN (LHC)~\cite{LHC} has started with a center-of-mass energy of $\sqrt{s}=7$~TeV
and
the LHC is expected to reveal the origin of the EWSB.
There are two possibilities, that is, scenarios with and without
the Higgs boson to describe the EWSB.
We focus on the latter scenario in this paper.
In case the Higgs boson does
not exist, there is no longer so-called naturalness problem.
However, in this case there are
problems in the unitarity of the longitudinal gauge boson scattering
\cite{Herrero:1998eq,Dicus:1992vj,Cornwall:1973tb,Cornwall:1974km,Lee:1977yc,Lee:1977eg,Chanowitz:1985hj}
and
in the consistency with electroweak precision tests.
During the past decade, models with extra dimensions have been studied as a new
paradigm. It has brought a solution of the gauge hierarchy, a solution
of the Yukawa hierarchy, many new particles called Kaluza-Klein (KK)
particles and dark-matter candidates (with the symmetry called KK parity).
The Higgsless model
\cite{Csaki:2003dt,Csaki:2003zu,Nomura:2003du,Barbieri:2003pr,Csaki:2003sh,Davoudiasl:2003me,Burdman:2003ya,Foadi:2003xa,Hirn:2004ze,Casalbuoni:2004id,Chivukula:2004pk,Csaki:2005vy,Csaki:2004sz}
is one of the extra-dimensional models. It does not
contain any physical scalar field. The EWSB is triggered by boundary
conditions for an extra-dimensional direction.
The Higgsless model can
keep a perturbativity,
and can be translated into a model in four dimensions
by the discretization of the extra-dimensional direction.
This translation is known as the deconstruction
\cite{Arkanihamed:2001ca,Hill:2000mu}.
The deconstruction allows us to interpret a gauge symmetry in
extra dimension as a direct product of infinite number of gauge
symmetries.
Hence a model with a direct product of finite number of gauge
symmetries can be
regarded as a low energy effective theory, where one of the ultraviolet
(UV) completions is described with
an extra-dimensional model.
Such models can be constructed by a bottom-up approach using non-linear
sigma models.
The structure
of their gauge sector
is based on a generalized hidden local symmetry \cite{Georgi:1989gp,Georgi:1989xy,Bando:1984ej,Bando:1985rf,Bando:1987ym,Bando:1987br,Harada:2003jx}.
The three-site Higgsless model \cite{SekharChivukula:2006cg}
is a minimal deconstructed Higgsless model
and contains three extra relatively heavy gauge bosons, $W^{\prime\pm}$ and $Z'$ as explained later.
The electroweak gauge symmetry of this model is $SU(2) \times SU(2) \times U(1)$,
which is broken down to $U(1)_{QED}$.
Hence this model is a low energy effective theory of the Higgsless model
in extra dimension and other UV complete models.
Through studies of this model,
we can find validity of other similar models at a time.
This is an advantage of this model,
and
it is important to find phenomenological constraints on this model.
Phenomenological constraints on this model are compatible with current
experiments \cite{SekharChivukula:2006cg,
Abe:2008hb,Abe:2009ni}.
Hence the LHC should be the most powerful experiment to test this model.
Because we have already know the precise bounds on parameters in this
model,
we can accurately predict physics and signals at the LHC.
There are papers on the three-site Higgsless model for the LHC~\cite{Ohl:2008ri,
Speckner:2010zi,
Han:2009em,Asano:2010ii,
Han:2009qr,Bian:2009kf,He:2007ge},
and
many of them are based on the so-called parton level analysis.
The parton level analysis ignores the effects of hadronizations and detector responses,
which are needed to get more realistic prospects.
Results, for example, the requirement of integrated luminosity
needed for new particle discovery, without considering such effects might be optimistic.
Therefore hadronizations and detector simulations are performed for the results shown in this paper.
According to the parton level analysis,
Drell-Yan (DY) production process of the $W^{\prime \pm}$ and $Z'$ bosons
is the most promising channel for their discovery.
Less integrated luminosity is required for the discovery of heavy gauge bosons
through the DY process than others.
As explained later,
the coupling among the $Z'$ and fermions, $g_{Z'ff}$,
is stronger than
the coupling among the $W'$ and fermions, $g_{W'ff}$ and
a parameter dependence of $g_{Z'ff}$ is more moderate compared with $g_{W'ff}$.
Therefore we focus on the DY production of $Z'$ in this paper,
and study it with a way beyond the parton level analysis.
This paper is organized as follows.
In section \ref{sec:3site-review},
we review the three-site Higgsless model briefly.
In section \ref{sec:analysis},
we perform feasibility studies with the experimental condition of the ATLAS experiment~\cite{ATLASDetector} for some signal points.
Section \ref{sec:summary}
is devoted for summary and discussion.
\section{Summary and Discussion}\label{sec:summary}
We have studied the discovery potential of the three-site Higgsless model
with the $Z'$ gauge boson via Drell-Yan production at proton-proton collision of a center-of-mass energy of $\sqrt{s}$ = 14~TeV.
The discovery potential of $Z'\rightarrow WW\rightarrow \ell\nu qq$ is evaluated with the simplified detector simulation of ATLAS experiment condition
which takes into account the effect of hadronization and experimental efficiency and resolution.
The significance is obtained with the event counting in the signal mass window.
We show that the significance reaches 3$\sigma$ threshold in the integrated luminosity 1-6 fb$^{-1}$~
and 5 $\sigma$ threshold in the 3-20 fb$^{-1}$~ for theoretically preferable $Z'$ mass region, 380-610 GeV.
The required integrated luminosity for the observation corresponds to a few years' running of the LHC at $\sqrt{s}$ = 14~TeV.
Depending on the model parameters,
we have found less integrated luminosity is required for the discovery
of the $Z'$ boson through the Drell-Yan process than other production processes.
On the other hand,
for the discovery of the $W^{\prime \pm}$ boson,
other production processes have advantage rather than the
Drell-Yan process
because of the fermiophobity of the $W^{\prime \pm}$ boson.
It is important to discover the $W^{\prime \pm}$ boson
because we have to check the origin of the $Z'$ boson is the $SU(2)$
gauge symmetry for a verification of Higgsless models.
There are many models which predict
the $Z'$ boson
which is not assosiated with $SU(2)$ gauge symmetry\cite{Langacker:2008yv}.
Precision predictions in such $Z'$ production are studied
in\cite{Fuks:2007gk,Fuks:2008mi,Rizzo:2006nw,Dittmar:2003ir}, for example.
Hence we need to discover $W^{\prime \pm}$ for a verification of Higgsless
models.
There is a mutually complementary relationship between the Drell-Yan
process and other production processes
in order to verify Higgsless models at the LHC.
|
\section{The topological order parameters}
The two main methods to calculate a topological order parameter for mixed states are the topological entropy of Levin and Wen \cite{levwen} (and a variant by Castelnovo and Chamon \cite{claudios1}) and the topological mutual information of Iblisdir et al. \cite{iblisdir} (based on the pure state topological entanglement entropy of Kitaev and Preskill \cite{kitpres}). Both of these measures result in the same value of $2\log D$ for topologically ordered states associated with an anyon model of total quantum dimension $D$.
The method of Levin and Wen notes that states of topologically ordered systems have correlations for regions that form closed loops that do not exist for those along open loops \cite{levwen}. As such an entropic difference between the two cases is expected. This can be measured by taking an annulus and dividing it into two regions, $A$ and $B$, then further dividing $B$ into $B_1$ and $B_2$, as shown in Fig. \ref{fig1}(a). Using these,
\begin{eqnarray} \label{LW0}
\gamma_{LW} + \ldots &=& (S_{B} - S_{B_1})-(S_{AB} - S_{AB_1}).
\end{eqnarray}
Here the ellipsis represents contributions from non-topological correlations, which will be suppressed as the regions are made larger than the correlation length. In this work, we note that this calculation of the topological entropy may alternatively be expressed in terms of mutual informations as,
\begin{equation} \label{LW}
\gamma_{LW} + \ldots = I_{A,B} - I_{A,B_1}.
\end{equation}
Note that this is not a modification of the quantity, just an alternative way of expressing it. By expressing it in this way we see that it is the information that one section of a closed string ($A$) shares with the rest ($B$), but does not share with its neighbouring sections ($B_1$). This is because $A$ is correlated to $B$ by both topological and local correlations, but is correlated to $B_1$ by only local correlations. The difference between these mutual informations thus isolates the topological correlations. There is no reason to suppose that this argument applies only to pure states. The entropy as defined by Levin and Wen is therefore equally applicable to mixed states. As such, this entropy has been applied to mixed states in previous studies \cite{claudios1,claudios2,melko}. It should be noted that a modified definition for mixed states has also been proposed in \cite{claudios1}, but it gives equivalent results and so will not be considered explicitly here.
The topological mutual information of Iblisdir et al. is a generalization of the topological entanglement entropy of Kitaev and Preskill \cite{kitpres} to mixed states. It uses the fact that the mutual information between the region $R$ and its complement, $R_c$, will satisfy an area law of the following form,
\begin{equation} \nonumber
I_{R} = \alpha L -\gamma_{I} + \ldots.
\end{equation}
Here $\alpha$ is a positive constant and the correction $\gamma$ is due to topological correlations. The ellipsis represents contributions from non-topological correlations, which disappear as the regions are made larger than the correlation length.
To cancel out the boundary terms and isolate the topological term, the following linear combination of mutual informations is taken using the regions shown in Fig. \ref{fig1}(a),
\begin{eqnarray} \nonumber
\gamma_{I} + \ldots &=& - I_{A} - I_{B} - I_{C} - I_{ABC}\\ \label{KP}
&+& I_{AB} + I_{AC} + I_{BC}.
\end{eqnarray}
Note that the quantity called the topological mutual information in \cite{iblisdsir} is actually $-\gamma_I$. However, $\gamma_I$ is used here such that the values of both measures considered are positive for topologically ordered states. With this definition, topologically ordered states associated with an anyon model of total quantum dimension $D$ will have $\gamma_{LW} = \gamma_I = 2 \log 2$, and non-topologically ordered states will have $\gamma_{LW} = \gamma_I = 0$.
\begin{figure}[t]
\begin{center}
{\includegraphics[width=8cm]{fig1.eps}}
\caption{\label{fig1} The regions used for the calculation of a: (a) the topological entropy in Eq. (\ref{LW}); (b) the topological mutual information in Eq. (\ref{KP}). For both of these entropy calculations, the size of the regions should be taken to be large for the most accurate results.}
\end{center}
\end{figure}
It is important to note that $\gamma_{LW}$ and $\gamma_I$ are measures of topological correlations, not of entanglement. For pure states, the distinction is unnecessary, since the presence of topological (or any other) correlations implies that entanglement is present. However, for mixed states, topological correlations may arise from classical probability distributions, as noted previously in \cite{claudios1}. This property can cause some ambiguity, since it is possible to confuse `true' topological order which is quantum in nature with classical topological order. However, it is not this problem that we address here.
\section{Long-range non-topological correlations}
The corrections present in the calculations above are due to non-topological correlations that may be present in the system. These can be characterized by by two-point correlation functions of the form $G_{i,j} = \langle O_i O_j \rangle - \langle O_i \rangle \langle O_j \rangle$, where $O_i$ is an operator defined in the neighbourhood of a point $i$ of the system. It is known that, for (pure) ground states of gapped Hamiltonians, $G_{i,j} = O(e^{-d/\xi})$, where $d$ is the distance between points $i$ and $j$. As such, these correlations are suppressed beyond the finite correlation length $\xi$. Making all length scales for the regions used in the above calculations much larger than $\xi$ therefore successfully suppresses the corrections. However, difficulties can arise when mixed states are considered. Though $G_{i,j}$ decays with distance for each individual degenerate ground state of a gapped Hamiltonian, it does not necessarily do so for a mixture of them. The mixed state can therefore, effectively, have an infinite correlation length. Such long-range non-topological correlations will then cause the corrections to remain finite for regions of arbitrary size, and the results of the calculations in Eq. (\ref{LW}) and Eq. (\ref{KP}) to give an ambiguous characterization of topological order.
As we will show below, this effect is not restricted to unrealistic states. Instead, it can occur for ground states of local gapped Hamiltonians. Furthermore, since the equally weighted mixture of degenerate ground states describes the thermal state at zero temperature, it is an important and realistic state to consider for any Hamiltonian. As such, it is vitally important to determine a definition of the topological entropy that is immune to such effects.
\begin{figure}[t]
\begin{center}
{\includegraphics[width=4cm]{fig2.eps}}
\caption{\label{fig2} The spin lattice on which the toric code, and hence the states $\rho^{(1)}$ and $\rho^{(2)}$, is defined. Dots denote spins, and operators are defined around each plaquette and vertex. The lattice has periodic boundary conditions.}
\end{center}
\end{figure}
Two simple examples of states with long-range non-topological correlations can be defined using the toric code model \cite{dennis} with a modified Hamiltonian. This model uses the spin lattice of Fig. \ref{fig2}, defining operators $A_v = \prod_{i \in v} \sigma^x_i$ on the spins around each vertex $v$, and operators $B_p = \prod_{i \in p} \sigma^z_i$ around the spins of each plaquette $p$. The usual Hamiltonian of the model is,
\begin{equation} \nonumber
H_0 = -J \sum_v A_v -J \sum_p B_p.
\end{equation}
The $A_v$ and $B_p$ operators define the occupation of anyons. States within their $-1$ ($+1$) eigenspace denote that an anyon is present (not present) on the corresponding vertex or plaquette. The Hamiltonian assigns energy to anyons, and hence the ground states of the Hamiltonian are those with no anyons anywhere. Any eigenstate of the Hamiltonian is topologically ordered, with $\gamma_{LW} = \gamma_I = 2 \log 2$. The same is true for any mixture of eigenstates if they all correspond to the same anyon configuration. As such, the ground states of the Hamiltonian have $\gamma_{LW} = \gamma_I = 2 \log 2$, as does any mixture of them. We will use $\rho^{(0)}$ to denote their equally weighted mixture.
For mixtures of $H_0$ eigenstates with different anyon configurations, the topological order parameters can take any value from $0$ to $2 \ln 2$, inclusive. The value $\gamma_{LW} = \gamma_I = 2 \log 2$ is only obtained if the following condition is met \cite{hastings}. Consider an annulus, such as that formed by the region $AB$ in Fig. \ref{fig1} (a). $\gamma_{LW} = \gamma_I = 2 \log 2$ if, by local operations on the spins within the annulus, a state can be prepared such that the parity of anyons on both the plaquettes and vertices enclosed by the the annulus (i.e. those with either full or partial support on spins within the region enclosed by the annulus) is in a definite state.
Let us now consider the following modified version of the toric code Hamiltonian,
\begin{equation} \nonumber
H_1 = -J \sum_{\langle v,v' \rangle} A_v A_{v'} -J \sum_{\langle p,p' \rangle} B_p B_{p'}.
\end{equation}
Here $\langle v,v' \rangle$ denotes a pair of vertices joined by an edge, and $\langle p,p' \rangle$ denotes a pair of plaquettes which share an edge. Hence, rather than single vertex and single plaquette terms of $H_0$, the Hamiltonian is composed of nearest neighbour two-vertex and two-plaquette interactions. Nevertheless, it is important to note that it is still both gapped and local, and has the same eigenstates as $H_0$. The ground states of $H_1$ correspond to the states for which the either no vertices hold an anyon, or all vertices hold an anyon. The same is true for the plaquettes. Let us consider the equally weighted mixture of all ground states, $\rho^{(1)}$.
In this case, the natural choice for the operator $O_i$ for the correlation function $G_{i,j}$ is $A_v$ (alternatively, $B_p$). It can then be easily seen that $G_{v,v'}(\rho^{(1)}) = 1$ for any pair of vertices $v$ and $v'$, without any decay due to the distance between them. The state therefore effectively has an infinite correlation length, despite being composed of ground states of a gapped and local Hamiltonian.
It is clear that this state is topologically ordered, with $\gamma_{LW} = \gamma_I = 2 \log 2$, since measurement of even a single vertex and plaquette on an annulus is sufficient to determine the occupancies of all vertices and plaquettes. The parity of both vertex and plaquette occupations enclosed by the annulus is then both definite and known.
To perform the calculation of the topological mutual information, note that state $\rho^{(1)}$ can be decomposed as an equally weighted mixture of four mixed states. Each of these corresponds to a different anyon configuration (vertices and plaquettes all empty, vertices filled and plaquettes empty, vertices empty and plaquettes filled, and vertices and plaquettes all filled), and hence they are orthogonal. Also, the entropies for each of these mixed states are equal, since they are equivalent by local unitaries. The reduced density matrices $\rho^{(1)}_R$ and $\rho^{(1)}_{R_c}$ are similarly composed of four orthogonal and locally unitarily equivalent mixed states, since the different anyon configurations can also be distinguished and manipulated within each region. The behaviour of the Von Neumann entropy when applied to block diagonal matrices \cite{iblisdir} then means that the mutual information can be decomposed into two contributions. The first comes from the correlations within each of the orthogonal and locally equivalent states. Since these states are all equivalent to $\rho^{(0)}$, this contribution is $I_{R}(\rho^{(0)})$. The second contribution comes from the fact that there are four states, distinguishable in both regions, that are equally mixed. This adds an amount $2 \log 2$ to the total mutual information. Hence,
\begin{equation} \label{decomp}
I_{R}(\rho^{(1)}) = I_{R,R_c}(\rho^{(0)}) + 2 \log 2.
\end{equation}
When these mutual informations are combined according to Eq. (\ref{KP}), the contribution of $I_{R}(\rho^{(0)})$ from each will lead to a total contribution of $2 \log 2$. This is due to the topological correlations within the state $\rho^{(0)}$. However, the total contribution of the $2 \log 2$ terms will be $-2 \log 2$. In total, therefore, the calculation for the topological mutual information yields $\gamma_I+\ldots=0$.
This result may lead anyone calculating this quantity to believe that $\rho^{(1)}$ does not possess topological correlations, and therefore that the nearest neighbour interactions of $H_1$ are not sufficient to support topologically ordered states. However, such a conclusion would be incorrect. The result of $\gamma+\ldots=0$ arises in this case only because non-topological correlations are not suppressed, no matter how large the regions are made, and are therefore able to cause ambiguity.
In general, any state on a 2D surface for which the spins (or other physical medium) around one point have correlations with other those around points spread throughout the surface, and the correlations do not decay with distance, will cause ambiguities in the calculation of the topological mutual information. If pure states occur with the same properties, similar ambiguities will arise for the topological entanglement entropy that it is based upon \cite{kitpres}. This is because such correlations cause the mutual information across any bipartition to have contributions that do not depend on the length of the boundary, and hence cannot be distinguished from genuine topological contributions.
For the second example we consider another modified toric code Hamiltonian,
\begin{equation} \nonumber
H_2 = -J \sum_{( v,v' )} A_v A_{v'} -J \sum_{( p,p' )} B_p B_{p'}.
\end{equation}
Here $(v,v')$ denotes a pair of vertices joined by a vertical edge, and $(p,p')$ denotes a pair of plaquettes which share a horizontal edge. As such, $H_2$ is like $H_1$ in that it consists of nearest neighbour 2-vertex and 2-plaquette interactions. But, unlike $H_1$, these interactions are confined to a specific direction. The ground state of $H_2$ consist of any state for which each vertical line of vertices or plaquettes is either empty or full of anyons. We consider the equally weighted state $\rho^{(2)}$ of all ground states.
As before, using $O_i = A_v$ yields $G_{v,v'}(\rho^{(2)}) = 1$ for any pair of vertices $v$ and $v'$ on the same line. There is again no decay due to distance, and hence an infinite correlation length despite being composed of ground states of the a gapped and local Hamiltonian.
This state is again topologically ordered with $\gamma_{LW} = \gamma_I = 2 \log 2$. This is because measurement of a vertex (plaquette) on each line of vertices (plaquettes) that intersect the region enclosed by the annulus is sufficient to determine the parity of the vertex (plaquette) occupations within this region.
Like $\rho^{(1)}$, the state $\rho^{(2)}$ is a sum of orthogonal and locally equivalent states of different anyon configurations, with each state locally equivalent to $\rho^{(0)}$. The same is true for the reduced states for any region. The entropy of a region $R$ is therefore composed of an $S_R(\rho^{(0)})$ contribution due to the correlations within each state and a contribution of $\log 2$ for each vertical line of plaquettes or vertices that intersects the region (i.e. has at least one vertex or plaquette with full support on the region). If we use $n_R$ to denote the number of such lines, the mutual information between two regions is
\begin{equation} \label{decomp}
I_{R,R'}(\rho^{(2)}) = I_{R,R'}(\rho^{(0)}) + (n_R + n_{R'} - n_{RR'}) \log 2.
\end{equation}
When these are combined according to Eq.(\ref{LW}), the result will be $\gamma_{LW}+\ldots= 2 \log 2 + n \log 2$. Here $n$ denotes the number of vertical lines that intersect both $A$ and $B_2$ but do not intersect $B_1$. The value of the topological entropy is therefore overestimated by an amount that is not suppressed as the size of the regions are increased, but in fact increases. This is due to the non-topological correlations along lines of anyons. The same is true for the reformulation of the Levin and Wen method in \cite{claudios1}.
In general, any states with infinite range non-topological correlations along lines are able to cause ambiguities to the topological entropy of Levin and Wen. This is because such correlations are able to directly correlate the regions $A$ and $B_2$, and the calculation is built on the assumption that such correlations do not occur.
With these examples, we see that both of these topological order parameters are susceptible to long range non-topological correlations, since they can cause corrections that cannot be suppressed. This may then lead to non-topological phases being identified as topological by mistake, or vice versa, especially as more exotic systems are probed \cite{zang,zheng,sen,son}. Though both examples considered only fool one of the parameters, with $\rho^{(1)}$ and $\rho^{(2)}$ both giving the expected values of $2 \ln 2$ when using $\gamma_{LW}$ and $\gamma_I$ respectively, clearly states will exist which fool them both. As such, a way to rid the entropy of the non-topological corrections must be determined.
\section{Modifying the topological entropy}
To solve the problem, let us take a closer look at how the topological entropy is defined. In Eq. (\ref{LW}) it was shown that $\gamma_{LW}$ is the information that a section of a closed string ($A$) shares with the rest of the string ($B$), but not with the neighbouring sections ($B_1$). Since it can usually be expected that any contributions to $I_{A,B}$ due to local correlations act over the boundary between $A$ and its neighbour $B_1$, this definition serves to remove all local boundary correlations and thus, in most cases, measures only the topological correlations around the string. The problem, as mentioned above, is that sufficiently long range local correlations contribute in a way that does not depend on the boundary, and can correlate the non-neighbouring regions $A$ and $B_2$ directly. Solving the problem therefore requires modifying the calculation of the topological entropy such that local correlations do not cause unsuppressable corrections, whatever their range. To do this let us define the topological entropy $\gamma$ as the information that the region $A$ shares with $B$, but not $B_1$ or $B_2$ individually.
This definition isolates the correlations that can be seen only when all the three regions of the annulus are considered. Only correlations that contribute to $I_{A,B}$, but not to $I_{A,B_1}$, $I_{A,B_2}$ or $I_{B_1,B_2}$ will contribute to $\gamma$. Since non-topological correlations characterized by two-point correlation function will indeed contribute to the latter three mutual informations, they cannot affect the value of $\gamma$, and so cannot cause ambiguity whatever their range. Only truly non-local correlations, such as those existing along strings in topological ordered states, will therefore remain. Note that this definition only requires that the three regions together form an annulus, and the composite of any two regions does not. As such partitioning the closed loop according to Fig. \ref{fig1}(b) is therefore no longer a requirement (though it is still valid, and we will continue to use in this study), and other tripartions of an annulus may be used.
The only caveat is as follows. Suppose that infinite range non-topological correlations are present in a state such that the operators $O_i$ are not defined on a single spin $i$, but of spins within its neighbourhood also. Consider then a point $j$ that lies on the boundary between the regions $B_1$ and $B_2$, such that the support of $O_j$ is split between these two regions, and consider also a point $i$ that lies fully within the region $A$. Any two point correlations between $i$ and $j$ can therefore only be revealed when all three regions are considered. Such correlations would therefore appear to contribute to $I_{A,B}$ without contributing to $I_{A,B_1}$, $I_{A,B_2}$ or $I_{B_1,B_2}$, and so cause ambiguities with the value of $\gamma$. However, assuming the Hamiltonian causing the correlations is local, the points $i$ and $j$ will not be correlated in isolation. There will be other points $k$ with which $i$ shares exactly the same information it shares with $j$. In the large annulus limit, at least some of these will exist such that that $O_k$ has full support within $B_1$ or $B_2$. The contribution to $I_{A,B}$ will also be present in $I_{A,B_1}$ or $I_{A,B_2}$, respectively, and so be removed from the calculation along with all other non-topological contributions.
Calculation of the modified topological entropy is not trivial, even when only classically correlated states are considered. Any calculation must distinguish the information that $A$ shares with both $B_1$ and $B_2$, and that which it shares only with the composite region $B$. In order to make the distinction, and calculate the modified topological entropy, we can use the state of the system, $\rho$ to define an alternative state $\rho'$. This is constructed such that $\rho_{AB_1} = \rho'_{AB_1}$ and $\rho_{AB_2} = \rho'_{AB_2}$. The correlations between the regions $A$ and $B_1$ are therefore exactly the same for $\rho'$ as they are for $\rho$, as are the correlations between the regions $A$ and $B_2$. We then maximize the entropy $S(\rho'_{AB})$ over all states for which these conditions hold. Any additional correlations that $A$ has only with the composite region $B$ will then not occur in $\rho'$. The topological entropy is then,
\begin{equation}
\gamma = I_{A,B}(\rho)-I_{A,B}(\rho').
\end{equation}
This calculation involves both reconstruction of a joint state from reduced density operators and a maximization. As such, computation of this quantity may be difficult in general. A simple method of calculating $\gamma$ in this way does exist if the states are classical, i.e. $\rho$ takes the form $\rho = \sum_{a,b_1,b_2} p(a,b_1,b_2) \ket{a,b_1,b_2}\bra{a,b_1,b_2}$, where $\{\ket{a}\}$ is a basis for the states of $A$, etc. In this case the state $\rho'$ can then be defined using the probabilities $p'(a,b_1,b_2) = p(a)p(b_1|a) p(b_2|a)$, the product of the marginal distributions for $B_1$ and $B_2$ conditioned on $A$. This maintains the correlations between $A$ and $B_1$, and between $A$ and $B_2$, but breaks those between $A$ and $B$ alone.
Even without an efficient means to calculate $ \gamma$ in general, both upper and lower bounds may easily be determined, applicable in both the quantum and classical cases. To calculate these bounds, we consider the quantity $I_{A,B}- \gamma$, the information $A$ shares with $B_1$ and $B_2$ individually but not with $B$ alone. This satisfies the bounds,
\begin{eqnarray}\nonumber
I_{A,B}- \gamma &\geq& \max(I_{A,B_1},I_{A,B_2}), \\ \nonumber
I_{A,B}- \gamma &\leq& I_{A,B_1}+I_{A,B_2}.
\end{eqnarray}
The first of these come from the fact that $I_{A,B}- \gamma$ is at least the information $A$ shares with $B_1$, or that which it shares with $B_2$. Taking the maximum of the two hence forms a lower bound. The second comes from the fact that adding the information $A$ shares with $B_1$ to that which it shares with $B_2$ will count all of the information in $I_{A,B}- \gamma$, and may even double count some, and so forms an upper bound. Using these, and noting that $\gamma$ cannot be less than zero, it follows that,
\begin{eqnarray} \label{new} \nonumber
\gamma &\geq& \min(I_{A,B} -I_{A,B_1}-I_{A,B_2},0), \\
\gamma &\leq& I_{A,B} - \max(I_{A,B_1},I_{A,B_2}).
\end{eqnarray}
This gives bounds for $\gamma$ in terms of the mutual informations for bipartitions. Note that these are exactly the kinds of quantities used to evaluate Eq. (\ref{KP}) and Eq. (\ref{LW}) above. As such, computation of these bounds is no harder than computation of existing quantities.
To demonstrate that the modified topological entropy correctly identifies topologically ordered states, the bounds may be calculated for the string-net models \cite{stringnet}, using the relations from \cite{levwen}. For these, the entropy of a region $R$ is given by,
\begin{equation} \label{topo}
S_R = - j_R \log D - n_{R} \sum_a \frac{d_a^2}{D} \log \left( \frac{d_a}{D} \right),
\end{equation}
where $j_R$ is the number of disconnected boundary curves in $R$ and $n_{R}$ is the number of spins along the boundary (see \cite{levwen} for more details on the specific models considered). The two regions $A$ and $B_2$ are not neighbouring, and so $j_{AB_2'}=j_A+j_{B_2}$ and $n_{AB_2}=n_A+n_{B_2}$. As such, we find that $I_{A,B_2}=0$, and so the regions are not correlated. The upper and lower bounds of Eq. (\ref{new}) then coincide both with each other, and with the definition of $\gamma$ in Eq. (\ref{LW}). This gives the topological entropy a unique value for these states, equal to the expected result of $ \gamma=2\log D$.
For the state $\rho^{(1)}$, $I_{R,R'}(\rho^{(1)}) = I_{R,R'}(\rho^{(0)}) + 2 \log 2$. As such, since we know that the topological correlations result in $I_{A,B}(\rho^{(0)}) - I_{A,B_1}(\rho^{(0)}) = 2 \log 2$, we see that the lower bound of Eq.(\ref{new}) vanishes and the upper bound gives $2 \log 2$. For the state $\rho^{(2)}$, the lower bound vanishes and the upper bound diverges with increasing region size.
The bounds for $\rho^{(1)}$ and $\rho^{(2)}$ therefore do not assign a unique value to $\gamma$ for in either case, nor can they tell us whether $\gamma=0$ or $\gamma>0$. Nevertheless, the fact that they are not equal is a clear signature of long-range non-topological correlations. Once their presence has been detected, efforts to remove the ambiguity they cause can be made. This is in contrast to previous methods, which assign a single value and give no clue as to whether this stems from topological correlations, or long range non-topological correlations.
\section{Conclusions}
Here we have shown that existing definitions of the topological entropy and topological mutual information can give misleading results, since they cannot completely distinguish between topological and non-topological correlations in all cases. As such, a modified definition of the topological entropy, $ \gamma$, has been proposed. This is equivalent to the topological entropy of Levin and Wen in most cases. However, unlike existing definitions, $ \gamma$ is robust against all non-topological correlations, even if they effectively have an infinite correlation length. Upper and lower bounds on this modified entropy have been determined, which can be computed with the same complexity required for previous methods. These bounds will coincide in most cases and so provide a unique value for $ \gamma$. In the the presence of long range non-topological correlations, the bounds diverge. This serves as a witness to the presence of such correlations, as well as limiting the ambiguity they can cause.
It should be noted that problems arising from infinite correlation lengths do not occur only for mixed states. Pure eigenstates of gapless Hamiltonians will similarly cause ambiguities in the calculation of the topological entropy. Though a full treatment of gapless states and the ambiguities they cause to the topological entropy is beyond the scope of this work, it seems likely that the modified topological entropy defined here will be useful also in the gapless case.
\section{Acknowledgements}
The author would like to thank Alioscia Hamma for critical reading of the manuscript and useful discussions, and especially for pointing out that non-topological correlations have a cancellation effect on $\gamma_I$. Thanks also to Jiannis K. Pachos and Abbas Al-Shimary for critical reading of the manuscript and useful discussions.
|
\section{Introduction}
The comparison of hadronic collider (\pp~collisions) and electron
collider (\ee~collisions) data have shown a clear discrepancy in
the produced particle multiplicity for collisions measured at the
same center-of-mass
energy. This has long been explained by the observation that
hadronic interactions can be decomposed in terms of ``collision
products'' and ``leading hadrons'' -- with the latter carrying away
(on average) half of the total energy available for particle
production~\cite{cite:LeadingHadrons}. It is found that one can
unify these data by shifting the center-of-mass energy of the
\pp~collisions to an effective center-of-mass energy, which is half
of the original collision energy, empirically determined.
Superimposing central heavy ion data onto this agrees
more closely with the \ee~data than the hadronic collision (\pp)
data~\cite{cite:PHOBOS_Universality}. Combining all these
observations, one can surmise that collisions in which all
``constituents'' of the proton are used (i.e. no leading hadron) will
produce higher (\ee-like) multiplicities.
The model presented here, originally described in
Ref.~\cite{cite:RichardsThesis}, uses these ideas as a premise
to divide the
participant region in heavy ion collisions into two distinct
regions; these are noted as ``mono'' and ``multi''. A Glauber
model~\cite{cite:Glauber} is used to calculate the number of
participants and the number of times each nucleon is struck
by another from the opposing nucleus. Mono refers to nucleons
which undergo a single hit -- similar to \pp~interactions, which
presumably can ``liberate'' a leading hadron. Multi refers to
nucleons which are multiply struck (by two or more) nucleons,
however, these are only counted once and thus do not
accumulate as quickly as the number of binary collisions.
\begin{figure*}[ht]
\centering
\includegraphics[angle=0,width=0.95\textwidth]{DrawWhereCollisionsAre.eps}
\vspace{-20pt}
\caption{\label{fig:MonoMultiPositions}
(color online) Predicted relative positions of mono- (upper) and
multi-collisions (lower). The distributions are based on a Glauber
calculation at impact parameter $b$\,=2,\,5,\,8,\,12\,fm
(\npart\,$\sim$\,359,\,261,\,140,\,24, \nmono\,$\sim$\,28,\,38,\,33,\,12, and
\nmulti\,$\sim$\,331,\,223,\,107,\,12).
In this picture, the reaction plane is always along the $x$-axis.}
\end{figure*}
This phenomenological approach is similar to the successful
core/corona approach of EPoS~\cite{cite:CoreCorona}.
The main technical difference surrounds the expected position
with respect to the reaction plane, of the mono (or corona)
nucleons. The current model expects that the peak of mono
nucleons to be {\it along} the reaction plane (see
Fig.~\ref{fig:MonoMultiPositions}), whereas in the corona from EPoS
is expected to peak 90$^{0}$ away~\cite{cite:CoreCorona}. A more
recent `toy model' implementation by the EPoS
authors~\cite{cite:CoreCorona2}, however,
does use the same ``Glauber-method'' to determine core (multi) and corona
(mono) regions to describe the strangeness enhancement observed in
Au+Au and Cu+Cu
data. In that implementation no interactions between the core and
corona regions is assumed, each ``corona'' contributes one-half of
a minimum bias interaction. The authors note that fixing the mono
(corona) is not necessary. However, although they undergo different
assumptions, both can provide a good description of
the data. Where these models differ is in the extent to which
they can be (or at least have been) applied and the possible
origin of the two components, or regions, of the collision. By using
the model presented here, it is possible to not only describe with
success low-\pT~phenomenon observed in heavy ion collisions but also
high-\pT~phenomena. This can lead to hints as to the origin of
disappearing back-to-back jets and their reappearance at low-\pT.
\section{The Model}
The implementation of this model is very simple. A Glauber model
is used to count the number of times each nucleon is struck (the
Glauber model implementation is similar to
Ref.~\cite{cite:PHOBOS_Glauber}). For
each Monte-Carlo ``collision'' the nucleons are sorted into three
categories: not hit (spectators), singly hit nucleons (\nmono), and
multiply hit nucleons (\nmulti). Spectators are not considered in
this model. After many trials, and at all possible impact parameters,
the average number of mono- and multi-assigned nucleons are
calculated. Figure~\ref{fig:MonoMultiVsNpart} illustrates the mean
\nmono~and \nmulti~found for Au+Au (black circles) and Cu+Cu (grey)
collisions at \snn\,=\,200\,GeV. \nmono~and \nmulti~versus centrality
are not found to appreciably change with collision energy, as shown
by lines depicting \nmono~and \nmulti~for 19.6\,GeV collisions.
After calculating the number of mono and multi participants for a
given centrality class it is possible to form the multiplicity
(and spectra, etc.) corresponding to that data using the formula:
\begin{equation}
\label{eqn:MonoMulti}
\ensuremath{ Y_{\rm Au+Au} = \nmono Y_{\rm mono} + \nmulti Y_{\rm multi}}
\end{equation}
\noindent $Y_{\rm Au+Au}$ represents the modeled Au+Au data, $Y_{\rm mono}$
is the yield expected from singly interacting nucleons (ostensibly
minimum bias \pp~data) and $Y_{\rm multi}$ is the yield expected from
the multi underlying distribution. $Y_{\rm mono}$ and $Y_{\rm multi}$
are fixed for a given data sample and do not change versus centrality.
$Y$ itself represents any distribution, for example multiplicity,
charged hadron spectra, kinetic freeze-out temperature, or any other
observable.
In this paper, we first explore the possible
origin of the underlying multi distribution which could be from one
of many different sources such as non-single diffractive (NSD) events
(unlikely as these may still have leading hadrons),
high-multiplicity events (for example greater than
$\langle\nch^{pp}\rangle$/2.), high-\pT~data ($\hat{p}_{T}$$>\sim$1.8
to represent the close-packedness of the multi nucleons), a
gluon-only source distribution. We use each of these different
hypotheses to test the systematic dependence -- although not to
choose a preferential particle production mechanism but rather to
provide a proof of principle that one can reasonably reproduce the
particle distributions in heavy ion collisions from these simple sources.
We note that a single distribution for mono (\pp~minimum bias) and a
single distribution for multi is assumed, which is sufficient to
reproduce all the data versus centrality.
The centrality dependence observed in the data is then entirely due to
the interplay between \nmono~and \nmulti, Fig.~\ref{fig:MonoMultiVsNpart},
and not from a changing or modified underlying distribution.
\begin{figure}[t]
\centering
\includegraphics[angle=0,width=0.475\textwidth]{drawBasic.eps}
\vspace{-20pt}
\caption{\label{fig:MonoMultiVsNpart}
The relative number of mono and multi collisions versus the number of
participants. Black (grey) circles represent Au+Au (Cu+Cu) collisions
at 200\,GeV ($\sigma_{NN}$\,=\,42\,mb); closed and open symbols represent
mono and multi collisions respectively. The values for 19.6\,GeV
($\sigma_{NN}$\,=\,33\,mb) Au+Au data are depicted as lines for comparison.}
\end{figure}
Secondly, as a more rigorous test, fits to the data to extract the underlying mono
and multi distributions will be made. In fact this is necessary
to analyze more exotic data than multiplicity and charged hadron
spectra. In this way we decouple ourselves from (for example)
\pythia~\cite{cite:PYTHIA} to describe the detailed identified
particle spectra. By using Eqn.~\ref{eqn:MonoMulti} and the
measured data to simultaneously extract the underlying
distributions, both the modification to the \pp~minimum bias
collisions at the surface (mono) as well as the true underlying
multi distribution can be studied.
As Glauber has no recourse to produce particle distributions, this
model has a purely phenomenological approach to describe the heavy
ion experimental data.
One important distinction to note for the current model versus many
other two-component models is the absence of ``collision scaling''.
Other two-component models assert that one component factorizes with
the number of participants (usually associated with soft particle
production), and the second with the number of nucleon-nucleon
interactions (\ncoll~-- representing hard collisions) within the
collision~\cite{cite:TwoComponentFit,cite:STARv2Model,cite:GyulassyWang,cite:KLN}.
In the current
discussion, we assert that the number of collisions is not the
correct scaling variable. If, as is hypothesized, upon
collision the system becomes a hot, dense QCD matter then it may become
difficult to reliably count individual nucleon-nucleon collisions.
The currently accepted approach used to validate the number of collisions
is to measure the spectrum of direct-photons in $A$+$A$ collisions,
relative to that in \pp~interactions. As the direct photons are not
expected to interact with the medium, it is assumed that the measured
\raa\,=\,1 validates the use of \ncoll~\cite{cite:PHENIX_dirgam},
illustrating that the number of originally produced photons are
consistent with \pp. In more current data, however, the previously
observed \ncoll~scaling at \pT\,$\sim$\,6\,GeV/$c$ is not seen at
higher momenta~\cite{cite:PHENIX_dirgamNew}, even though punch-through
high-\pT~hadrons are seen to escape the medium~\cite{cite:STAR_dijets}.
Here, we set this \ncoll~scaling assumption aside.
The number of mono and multi collisions in the current model is
less susceptible to the assumptions needed in order to count
the number of collisions, where our {\em number}
reflects an interaction volume (multi) and simply the number of
expected nucleon-nucleon scatterings on the surface (mono).
For a low number of participants (\npart$<$25), the number of mono
collisions exceeds
that of multi, and vice-versa for a high \npart. However, as shown
in Fig.~\ref{fig:MonoMultiVsNpart}, the number of mono does not fall to
zero even for the most central events where
approximately~7\% of the participants undergo only a single
interaction. An important consequence in these calculations is that
the sum of mono
and multi, by definition, is fixed to be \npart. This restricts the
growth of multi, unlike the rapidly growing number of binary collisions,
and results in multi (and mono) being more closely associated
to a ``participant'' variable. In calculating the mono and multi, it
is found that there is a strong dependence on the collision species,
in contrast to the number of binary collisions (\ncoll) which is
essentially the same at the same \npart. Similarly, only a weak
dependence on the collision energy (inelastic \pp~cross-section) is
observed over the energy range of the RHIC data, whereas a strong
binary collision dependence on \ncoll~is noted.
\section{Multiplicity Distributions}
As the premise of this model uses the charged particle multiplicity to
argue for distinct mono/multi regions of the collision volume it is
natural to begin with a discussion of that experimental data. Here,
one can simply use the measured \pp~distribution as the mono and a
second underlying distribution (for example taken from \pythia) to represent
the multi. We combine the mono/multi distributions (and similarly
for the next section) as:
\begin{equation}
\label{eqn:multiplicity}
\ensuremath{\frac{dN_{ch}^{\rm Au+Au}}{d\eta} =
\nmono\frac{dN_{ch}^{\rm mono}}{d\eta} +
\nmulti\frac{dN_{ch}^{\rm multi}}{d\eta}}
\end{equation}
\noindent where \nmono~and \nmulti~distributions for Au+Au
collisions are given in Fig.~\ref{fig:MonoMultiVsNpart}. One can
hypothesize that $dN_{ch}^{\rm mono}/d\eta$ be
$dN_{ch}^{\pp}/d\eta$ (i.e. precisely from minimum bias
\pp~interactions) or it can be a free parameter in the fit. For simplicity
in this first case, we assume the former; later, the latter is used to
test for any modification of the underlying mono distributions.
\begin{figure*}[th]
\centering
\includegraphics[angle=0,width=0.80\textwidth]{dNdeta_NoFit.eps}
\vspace{-24pt}
\caption{\label{fig:MultiplicityI}
(color online) Comparison between measured multiplicity versus $\eta$
(PHOBOS \snn\,=\,200\,GeV Au+Au data~\cite{cite:PHOBOS_dNdeta_200}) (band)
and model predictions with various candidates for the underlying
``multi'' $dN_{ch}^{\rm multi}/d\eta$~distribution as noted in the
legend. 0-6\% central (left) and mid-central
35-45\% (right) data are shown. In each model representation,
the ``mono'' is fixed to the inelastic \pp~distribution at the
same energy, i.e. $dN_{ch}^{\rm mono}/d\eta$\,=\,$dN_{ch}^{\pp}/d\eta$. Note the $y$-axis scale
difference in each figure.}
\end{figure*}
Figure~\ref{fig:MultiplicityI} shows the result of using this
formalism with several hypothesized multi multiplicity distributions
for Au+Au collisions at \snn\,=\,200\,GeV. For the shown centrality
bins from PHOBOS~\cite{cite:PHOBOS_dNdeta_200}, one can see that the model
provides a reasonable description of the data. Remember that the same
underlying distributions are used for both panels, only \nmono~and
\nmulti~are changed. This is not too
surprising as the basic premise of this model is to reproduce
the centrality dependence of the charged particle multiplicity in
the context of an absence of leading hadrons in the multi region.
From this, judging the model's success is really
only dependent on the choice of the underlying multi distribution.
Three simple \pp~multiplicity-like distributions are used for the
multi component, for comparison. The \pp~collisions at
\snn\,=\,200\,GeV minimum bias distribution is used (closed
circles) which is similar to the ``Wounded Nucleon
Model''~\cite{cite:WNM} approach and yields a multiplicity which is
$\sim$30\% lower than the measured Au+Au data. Rejecting the single
diffractive events yields a higher multiplicity (open circles)
but is still significantly lower than the Au+Au distributions. A more
radical approach is to only choose the highest multiplicity
\pp~events, in this case more than the average ($>$24) (squares),
which does achieve the desired multiplicity. One can
understand this approach in terms of the origins of the model --
as higher multiplicity events typically have lower energy leading
hadrons, thus releasing more energy to particle production.
Changing the reference \pp~multiplicity is not the only way to alter
the resultant multiplicity distribution. Also shown are a subset of
events from the minimum bias \pp~cross-section which exclusively
selects out gluon interactions ($gg\rightarrow gg$ or
$gg\rightarrow q\overline{q}$) with a large momentum transfer
($\hat{p}_{T}>1.8\pm0.1$\,GeV/$c$) (lines). (Note that the same multiplicity
enhancement is also found by using all (quark and gluon) interactions
with $\hat{p}_{T}>1.8$\,GeV/$c$.) Such an approach may represent
the close-packedness of the nucleons which may facilitate a large
momentum transfer. The different minimum $\hat{p}_{T}$ values
used as the multi component (denoted as dotted and dashed lines
in Fig.~\ref{fig:MultiplicityI}) show the sensitivity of the choice of
$\hat{p}_{T}$ in the final distribution
($dN_{ch}^{\rm Au+Au}/d\eta$ of Eqn.~\ref{eqn:multiplicity});
for multiplicity this is small.
Rather than assuming a given \ymono~and/or \ymulti~multiplicity
distribution,
Eqn.~\ref{eqn:MonoMulti} can be used to extract these two
underlying distributions from the data. Two bins in centrality
(preferably far apart in centrality) are used to simultaneously extract
\ymono~and \ymulti. This is referred to as the pair
fit method. After finding the underlying distributions, these are fixed
and all centrality bins are recombined using the \nmono~and \nmulti~as in
Eqn.~\ref{eqn:MonoMulti}.
To test the sensitivity to the bins chosen, several
combinations are made. The top panels of Fig.~\ref{fig:MultiplicityII}
show the PHOBOS data (grey bands) with the fit distributions (solid
lines). Panels (a,b), (c,d), (e,f), (g,h) in
Fig.~\ref{fig:MultiplicityII}, show data for Au+Au collisions at
200~\cite{cite:PHOBOS_dNdeta_200}, 130~\cite{cite:PHOBOS_dNdeta_130}, 62.4~\cite{cite:PHOBOS_62.4mid}, and 19.6\,GeV~\cite{cite:PHOBOS_CuCudNdeta} respectively. In each case, the model
uses the most central 0-6\% and 35-45\% (mid-central) data to
extract the underlying mono and multi distributions. Using other bin
combinations does not significantly alter the resultant distributions.
The lower panel of
Fig.~\ref{fig:MultiplicityII} shows the underlying \ymono~(dashed line) and
\ymulti~(solid line) distributions from the fit. The grey and open bands
represent the systematic error of the data. For
comparison, the gluon-only distribution selected from \pythia~(dot-dashed
lines for the $\hat{p}_{T}>1.8$\,GeV/$c$ distribution only at 200\,GeV)
along with the minimum bias at the same center-of-mass energy (circles) are also shown.
The underlying multi distribution is considered ``stable'' in the sense
that using different centrality pairs does not change appreciably the extracted
\ymulti~distribution. Mono, however, is less stable when using the pair fit method and is
found to be more dependent on the choice of centrality bin. Several
interesting artifacts of the fit are observed. First, the extracted
\ymulti~distribution is close to that which would be obtained
by using a gluon-only \pp~distribution in Eqn.~\ref{eqn:multiplicity},
but with more yield at mid-rapidity and also narrower.
Second, the mono distribution is considerably lower at mid-rapidity
than the measured minimum bias \pp~collisions~\cite{cite:ppRef200,cite:ppRef62.4_22.4}, as shown by the comparison of the
mono-collisions from the pair fit to minimum bias \pp~in
Fig.~\ref{fig:MultiplicityII}, panels (b), (f), and (h).
This is possibly hinting that
a modification exists, due to an opaque multi region.
A more detailed description of this modification is discussed in
the next section. The final notable observation is that the yield
of the extracted mono interactions is higher at forward rapidities,
as compared to the minimum bias \pp. This could possibly point to an
influence of spectators in the measured data.
To reiterate, \ymulti~represents the multiplicity per participant pair
of the multi region, which is consistent with the expected yield from
a gluon-dominated collisional center. The core is found to be opaque,
with some of the multiplicity from the surrounding mono collisions
absorbed. Importantly, even though the size of the central multi region
changes, the yield per participant does not change.
\begin{figure*}[ht]
\centering
\begin{minipage}{0.23\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_200.eps}
\end{minipage}
\begin{minipage}{0.23\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_130.eps}
\end{minipage}
\begin{minipage}{0.23\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_62.eps}
\end{minipage}
\begin{minipage}{0.23\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_19.eps}
\end{minipage}
\caption{\label{fig:MultiplicityII}
(color online) PHOBOS multiplicity data (bands) from Au+Au collisions
compared to
the multiplicity distributions derived from the model, lines, using the pair
fit method (upper panels). Panels (a,b), (c,d), (d,e), (f,g)
show data for Au+Au collisions at 200~\cite{cite:PHOBOS_dNdeta_200}, 130~\cite{cite:PHOBOS_dNdeta_130}, 62.4~\cite{cite:PHOBOS_62.4mid}, and 19.6\,GeV~\cite{cite:PHOBOS_CuCudNdeta}
respectively. The lower panels show the underlying \ymono~(dashed
line/open band) and \ymulti~(solid line/filled band). For reference,
the minimum bias \pp~data~\cite{cite:ppRef200,cite:ppRef62.4_22.4}
(green closed circles) and gluon only distribution (green dot-dashed
line) are shown (the latter for 200\,GeV only).
The open (closed) star symbols represent the result from the least-$\chi^{2}$ method
to determine the underlying mono (multi) from mid-rapidity data, see
Fig.~\ref{fig:MultiplicityIImid}.
}
\end{figure*}
\begin{figure*}[ht]
\centering
\vspace{-11pt}
\includegraphics[angle=0,width=0.375\textwidth]{dNdeta_midrap2.eps}
\includegraphics[angle=0,width=0.375\textwidth]{dNdeta_midrap2_CuCu.eps}
\vspace{-20pt}
\caption{\label{fig:MultiplicityIImid}
PHOBOS~\cite{cite:PHOBOS_200mid,cite:PHOBOS_130mid,cite:PHOBOS_62.4mid,cite:PHOBOS_19.6mid,cite:PHOBOS_CuCu_MidRap} and ALICE~\cite{cite:ALICE_2.76mid} multiplicity data at mid-rapidity compared to that
derived from the fit, using the least-$\chi^{2}$ method,
for Au+Au (Pb+Pb) data in the left panel and Cu+Cu data in the right
panel. Fits are shown for RHIC energies \snn\,=\,200\,GeV (lightest
grey band), 130\,GeV (light grey, Au+Au only)), 62.4\,GeV (dark grey), and
19.6\,GeV (22.4\,GeV for Cu+Cu) (darkest grey). The ALICE data (2.76\,TeV)
are shown as a no-fill outline. For comparison, in the right panel, the
expected Cu+Cu data are shown using the \ymono~and \ymulti~yields found
from the fits to the Au+Au collision data.}
\end{figure*}
\begin{figure*}[ht]
\centering
\includegraphics[width=0.75\textwidth]{MeanRadii.eps}
\vspace{-18pt}
\caption{\label{fig:MeanRadii}
(color online) Mean mono and multi radii -- estimated from the Glauber
model. The left (right) panel shows the radii along the ($90^{\circ}$ to)
reaction plane (the line along the nuclei centers -- $x$-
($y$-) axis in Fig.~\ref{fig:MonoMultiPositions}). The solid
line shows the multi radius (defined as 10\% of the peak).
The different dashed lines represent the mono peak position
(dash) and the outer/inner radius, defined as 10\% of the peak
yield. Negative radii (in the inner radius) represents the
point at which the mono are no longer distinctly separated by the
multi -- i.e. mono collisions occur at every point throughout the
collision area.}
\end{figure*}
\begin{table*}[th]
\caption{\label{tbl:MidRapVsEnergy} Mono and multi yields (and relative to
data \pp~inelastic
yields~\cite{cite:ppRef200,cite:PHOBOS_130mid,cite:ppRef62.4_22.4,cite:PHOBOS_19.6mid})
as derived from the least-$\chi^{2}$ fit to the
mid-rapidity data from PHOBOS (Au+Au and Cu+Cu) and ALICE (Pb+Pb).}
\begin{tabular}{| c || c | c || c | c || c | c || c | c |}
\hline
& \multicolumn{4}{c ||}{Au+Au} & \multicolumn{4}{ c |}{Cu+Cu} \\
Energy & \multicolumn{2}{c ||}{$dN_{ch}^{mono}/d\eta$} & \multicolumn{2}{ c ||}{$dN_{ch}^{multi}/d\eta$} & \multicolumn{2}{c ||}{$dN_{ch}^{mono}/d\eta$} & \multicolumn{2}{ c |}{$dN_{ch}^{multi}/d\eta$} \\
(GeV) & yield & yield/\pp & yield & yield/\pp & yield & yield/\pp & yield & yield/\pp \\\hline
2760 & 1.35 & -- & 8.55 & -- & -- & -- & -- & -- \\
200 & 1.49 & 0.65 & 4.01 & 1.75 & 2.07 & 0.91 & 3.83 & 1.67 \\
130 & 1.04 & 0.46 & 3.56 & 1.74 & -- & -- & -- & -- \\
62.4 & 0.86 & 0.45 & 2.88 & 1.50 & 1.58 & 0.82 & 2.75 & 1.43 \\
19.6 & 0.99 & 0.78 & 1.94 & 1.52 & 1.53 & 1.20 & 1.98 & 1.56 \\\hline
\end{tabular}
\end{table*}
There is another way to extract the \ymono~and \ymulti~yields from the data.
Figure~\ref{fig:MultiplicityIImid} (left panel) shows the mid-rapidity Au+Au data from
PHOBOS~\cite{cite:PHOBOS_200mid,cite:PHOBOS_130mid,cite:PHOBOS_62.4mid,cite:PHOBOS_19.6mid} and
Pb+Pb data from ALICE~\cite{cite:ALICE_2.76mid} data (bands) together with a least-$\chi^{2}$ fit (lines)
using the \nmono~and
\nmulti~distributions from Fig.~\ref{fig:MonoMultiVsNpart}. The right panel
of Fig.~\ref{fig:MultiplicityIImid} shows the Cu+Cu data from PHOBOS~\cite{cite:PHOBOS_CuCu_MidRap}
along with fits using that data and the extracted parameters from the fit to the Au+Au
collision data at the same energy. (Note that the 22.4\,GeV Au+Au reference is
scaled by 1.04 to account for the small difference in collision energy.)
This least-$\chi^{2}$ fit is a different approach than the pair fit method
used above. This enables the full centrality dependence of the mid-rapidity
results to be simultaneously utilized and reduces the dependence of the
results from the single pair choice. The
resultant least-$\chi^{2}$ fit at each energy well represents the data, even over
two orders of magnitude of collision energy. The derived mid-rapidity
mono and multi multiplicities
are summarized in Table~\ref{tbl:MidRapVsEnergy}.
For comparison,
these yields are shown as a star symbol to compared to the pair
fit method in Fig.~\ref{fig:MultiplicityII}.
In an attempt to reconcile the difference between the minimum bias
\pp~data and the derived underlying mono-distribution, we compare
the detailed spacial positions of the mono and multi interactions,
similar to those shown in Fig.~\ref{fig:MonoMultiPositions}.
Figure~\ref{fig:MeanRadii} shows the mean radius of the multi
(solid line) and mono (dot and dot-dashed lines) defined at
10\% of the peak. The center of the mono (peak) is shown as the
dashed line, which coincides with the multi radius. The
left (right) figures show the radii along ($90^{\circ}$ to) the
reaction plane. Clearly, half of the mono distribution resides
within the multi distribution. One could well imagine that if the
multi represents an opaque medium, then much of the mono inside
the multi region could be absorbed (or at least modified/suppressed).
Similarly, those outside the multi region may be suppressed should
the particles be headed toward the multi region, an eclipse-type
effect.
One can test the model by using the Cu+Cu
data~\cite{cite:PHOBOS_CuCudNdeta} to extract new \ymono~and \ymulti~yields
from the data and compare those to the Au+Au underlying
mono and multi distributions. Figure~\ref{fig:MultiplicityIICuCu}
shows a similar analysis as described above for the pair fit
method, this time with Cu+Cu collisions at \snn\,=\,200, 62.4, and
22.4\,GeV. For completeness, the underlying \ymono~and \ymulti~distributions
from Au+Au collisions (blue dashed lines) are also used to
form the multiplicity data.
In comparing the underlying mono and multi distributions from Cu+Cu
data to those from Au+Au data, we find that the multi distributions
are very similar and that the mono distribution in Cu+Cu more closely
resembles that from minimum bias \pp~interactions. Similarly in the
right panel of Fig.~\ref{fig:MultiplicityIImid}, the same
least-$\chi^{2}$ method used for Au+Au collisions is applied to the
mid-rapidity Cu+Cu data from PHOBOS~\cite{cite:PHOBOS_CuCu_MidRap} where
the same trends are observed in terms of the underlying distributions.
\begin{figure*}[t]
\centering
\begin{minipage}{0.24\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_200_CuCu.eps}
\end{minipage}
\begin{minipage}{0.24\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_62_CuCu.eps}
\end{minipage}
\begin{minipage}{0.24\textwidth}
\includegraphics[angle=0,width=1\textwidth]{dNdeta_PHOBOS_model_22_CuCu.eps}
\end{minipage}
\caption{\label{fig:MultiplicityIICuCu}
(color online) PHOBOS multiplicity data from Cu+Cu collisions~\cite{cite:PHOBOS_CuCudNdeta} compared
to the multiplicity distributions derived from the model, using the pair
fit method (upper panels, red) and from using the underlying \ymono~and
\ymulti~distributions from Au+Au collisions (blue, see
Fig.~\ref{fig:MultiplicityII}). Panels (a,b), (c,d), and (d,e) show
data for Cu+Cu collisions at 200, 62.4, and 22.4\,GeV respectively.
The lower panels show the underlying mono (dashed line/open band) and multi
(solid line/filled band) extracted from Cu+Cu data. The minimum bias
\pp~data~\cite{cite:ppRef200,cite:ppRef62.4_22.4} (green
circles) and
the underlying mono and multi distributions from the Au+Au analysis (blue dashed lines) are
shown for reference. The star (square) symbols represent the least-$\chi^{2}$
fit to the mid-rapidity Cu+Cu (Au+Au) data, see Fig.~\ref{fig:MultiplicityIImid} right (left) panel.}
\end{figure*}
\section{Unidentified Charged Hadron Spectra}
Following the success of the analysis of the charged particle
multiplicity, we apply this approach on more detailed data to check whether
a reasonable agreement could be found. First, the unidentified charged hadron
spectra versus transverse momentum (\pT) is used. This covers a
different kinematic region than the low-\pT~dominated multiplicity.
The suppression of high-\pT~particles in central events is a
fascinating phenomena which has been modeled by many different
theories, which are generally decoupled from the low-\pT~region.
Here, the full \pT~range of the data is used with the primary goal
of a qualitative
description of the data. As an initial test, the \ymono~and \ymulti~are
again taken from \pythia, at a collision energy of 200\,GeV, with
mono derived from a truly minimum bias sample and
multi from several hypothesized distributions as described above.
Again, \nmono~and \nmulti~are from Fig.~\ref{fig:MonoMultiVsNpart}.
Figure~\ref{fig:ChHadronSpectraI} shows the nuclear modification factor,
\raa~-- see Eqn.~\ref{eqn:raa}, for Au+Au collisions at \snn\,=\,200\,GeV
for two centrality bins as measured by PHENIX~\cite{cite:PHENIX_RAA_AuAu200}.
It can be clearly seen that,
although the model distributions do not match the experimental
data precisely, the qualitative features of the data are reproduced.
Using the \pp~minimum bias (circles) or high multiplicity (squares)
simulations for the multi component of the hadron spectra do not
represent the data enhancement
in the intermediate \pT~region with either of these as the underlying multi
distribution. The minimum bias result, in this case, represents a simple
visual scale of the number of collisions.
\begin{equation}
\label{eqn:raa}
\ensuremath{ \raa = \frac{1}{\ncoll}\frac{\frac{d^{2}N^{\rm Au+Au}}{dyd\pT}}{\frac{d^{2}N^{\pp}}{dyd\pT}}}
\end{equation}
\begin{equation}
\label{eqn:raanpart}
\ensuremath{ \raanpart = \frac{1}{\npart}\frac{\frac{d^{2}N^{\rm Au+Au}}{dyd\pT}}{\frac{d^{2}N^{\pp}}{dyd\pT}}}
\end{equation}
\begin{figure*}[h]
\centering
\includegraphics[angle=0,width=0.70\textwidth]{PhenixSpectra.eps}
\vspace{-24pt}
\caption{\label{fig:ChHadronSpectraI}
(color online) Comparison between measured nuclear modification factor,
\raa, versus \pT~(PHENIX data~\cite{cite:PHENIX_RAA_AuAu200}) (band)
and model predictions with various candidate \ymulti~distributions
(colored symbols) as noted in the legend. 0-10\% central (left) and
mid-central 40-50\% (right) data are shown. In each model
representation, \ymono~is fixed to minimum bias
\pp~from \pythia.}
\end{figure*}
Applying a minimum $\hat{p}_{T}$ cut of 1.8\,GeV/$c$ on the minimum
bias sample (dot-dashed line) results in a clear peak close to the
minimum cut (as expected) but over predicts the higher-\pT~data.
Removing all quark interactions, leaving a gluon dominated system
(solid, dotted, and dashed lines) both
reproduces the intermediate-\pT~peak and the lower yield at
higher-\pT. To emphasize again, the objective of this analysis is
not to {\it find} the underlying distribution, but to show that a
simple underlying distribution could exist which can describe the
features observed in heavy ion data.
\begin{figure*}[h]
\centering
\includegraphics[angle=0,width=0.90\textwidth]{newRAA_PHENIX.eps}
\caption{\label{fig:ChHadronSpectraII}
Nuclear modification factor (scaled by \npart~not \ncoll) for
unidentified charged hadrons from PHENIX~\cite{cite:PHENIX_RAA_AuAu200}
(filled band) at \snn\,=\,200\,GeV. The lines are the
\raanpart~reconstituted from the underlying mono- and multi-distributions
using the pair fit method; dashed (dotted) lines use centrality bins
0-10\% (20-30\%) and 60-70\%. In each figure, the dot-dashed line
represents the expected \ncoll~scaling, assuming particle production
scales as \pp~$\times$~\ncoll.}
\end{figure*}
Using the same technique to fit the data as described above for the
multiplicity `pair fit', charged hadron spectra can be modeled
using PHENIX data~\cite{cite:PHENIX_RAA_AuAu200}. The
fit is performed twice, once using the 0-10\% and 60-70\%
cross-section bins as the data pair to obtain the \ymono~and
\ymulti~distributions and a second pairing of 20-30\% and 60-70\% to test
the sensitivity in the choice of bin pairing, see
Fig.~\ref{fig:ChHadronSpectraII}.
The extracted underlying mono and multi distributions are then used
to model \raanpart, Eqn.\ref{eqn:raanpart}, in the remaining centrality bins.
\begin{figure*}[h]
\centering
\includegraphics[angle=0,width=0.375\textwidth]{newRAA_PHENIX_monomulti.eps}
\includegraphics[angle=0,width=0.375\textwidth]{PHENIX_MonoMulti_Spectra.eps}
\vspace{-20pt}
\caption{\label{fig:ChHadronSpectraIIresult}
(color online) Mono and multi nuclear modification factors (scaled by
\npart~not \ncoll) using the pair fit method on PHENIX,
PHOBOS~\cite{cite:PHOBOS_RAA_AuAu200}, and
STAR~\cite{cite:STAR_RAA_AuAu200} data. The left panel shows
\snn\,=\,200\,GeV mono and multi distributions,
derived from a fit to the PHENIX data. The dashed and dotted lines
show the fits from the centrality bins described in
Fig.~\ref{fig:ChHadronSpectraII}; the outline shows the 1-$\sigma$
band, derived from all possible centrality bin pairs. The grey band
shows equivalent PHENIX $d$+Au data~\cite{cite:PHENIX_dAu_Spectra}
from minimum bias collisions for reference. The closed (open) circles
represent the results from the least-$\chi^{2}$ method applied to the
spectra (omitting the four lowest centrality points from the fit).
The right panel shows the mono and multi spectra using the least-$\chi^{2}$
fit method. The closed red (open blue) circles represent the multi
(mono) collisions spectra per participant. The solid line represents
the 200\,GeV pp spectra, taken from~\cite{cite:PHENIX_RAA_AuAu200}.}
\end{figure*}
\begin{figure}[h]
\centering
\includegraphics[angle=0,width=0.375\textwidth]{newRAA_compExpt.eps}
\caption{\label{fig:ChHadronSpectraIIresultComp}
(color online) Comparison of the mean mono and multi distributions
derived from fits to PHOBOS (solid lines), PHENIX
(dashed), and STAR (dotted) for Au+Au collisions at 200\,GeV.}
\end{figure}
We find that all multi-dominated centrality bins are well reproduced
from the underlying mono and multi distributions using any two bins;
the results are not sensitive to the choice of centrality bins. In
each figure, the double-dot dashed line represents the `\ncoll'
scaling. For data presented as
\raa, the line would be exactly at unity and the data yield (at
high-\pT) would represent the apparent suppression. Within the mono/multi framework, we find that suppression at
high-\pT~is not necessary to describe the data. The most
peripheral bins, perhaps, are not as well described. An explanation
for this limited agreement could be that a ``collectivity'' which
produces the multi distribution could be dissipating. This could
point to a possible need to modify the model to account for an
additional scaling variable for example mono, duo and multi$^{\ge 3}$;
this is outside the scope of the current paper.
\begin{figure*}[h]
\centering
\includegraphics[angle=0,width=0.90\textwidth]{RAA_PHOBOS_model_62.eps}
\caption{\label{fig:ChHadronSpectraII62}
Nuclear modification factor (scaled by \npart~not \ncoll) for
unidentified charged hadrons from PHOBOS~\cite{cite:PHOBOS_RAA_AuAu62}
(filled band) at \snn\,=\,62.4\,GeV. The lines are the
\raanpart~reconstituted from the underlying mono- and multi-distributions
using the pair fit method; solid (dashed) lines use centrality bins
0-6\% (15-25\%) and 45-50\%. In each figure, the dot-dashed line
represents the expected \ncoll~scaling, assuming particle production
scales as \pp~$\times$~\ncoll.}
\end{figure*}
Figure~\ref{fig:ChHadronSpectraIIresult} (left panel) shows the
underlying mono and multi nuclear modification factors for \snn\,=\,200\,GeV
derived from the pair fit method; with the corresponding spectra in the
right panel. The dashed and dotted lines
represent the centrality bin pairs used in the analysis shown in
Fig.~\ref{fig:ChHadronSpectraII}. The red outline shows a
1-$\sigma$ band around the mean underlying multi distribution
determined from an average over all possible centrality-pair
combinations, except the most peripheral. This appears to be
well constrained, illustrating only a small variance in the
underlying multi distribution. A comparative analysis using the
least-$\chi^{2}$ method (for the yields versus centrality for each
\pT~bin) was also performed. (Note that the lowest four centrality
points were omitted from the fit.) The results are shown as circles in
Fig.~\ref{fig:ChHadronSpectraIIresult} (left panel) and are in
good agreement with the distributions extracted from the pair fit
method. The extracted underlying mono distribution, however, is less
constrained, possibly due to a changing surface suppression of
mono with centrality. Nonetheless,
the extracted values of the mono distributions here are
in agreement with the ones obtained from fits to the Au+Au
charged particle multiplicity: at very low \pT, a suppression
is observed with respect to minimum bias \pp~collisions.
In the intermediate \pT~region the mono distribution falls to zero,
hinting that the absorption (or suppression) is maximal in this
kinematic region. For some fits, it is found that the suppression
is `negative' this would represent a case when some additional
suppression is present from the edge of the multi-region; pointing,
perhaps, toward the need for a mono, duo, and multi$^{\ge 3}$ prescription.
Figure~\ref{fig:ChHadronSpectraIIresultComp} shows a comparison of
the resultant distributions by using PHENIX,
PHOBOS~\cite{cite:PHOBOS_RAA_AuAu200}, and
STAR~\cite{cite:STAR_RAA_AuAu200} data. The multi
distribution in this analysis are found to be very similar.
To put the results into context, Fig.~\ref{fig:ChHadronSpectraIIresult}
shows the multi spectra (per participant) relative to the minimum bias
\pp~spectra, which is found to have the same systematic dependencies as
the gluon-dominated system found in Fig.~\ref{fig:ChHadronSpectraI}.
It has been seen that, through the separation into multi and mono
sub-components of the collision, a consistent picture can be observed.
At the center of this picture is a dense (perhaps gluon dominated) system
which is opaque to the particles produced at the periphery of the
collision. A gluon-dominated picture is not excluded by other
measurements, for example, the anomalous increase of baryons, with
respect to mesons, could be expected from a system with a higher number
of gluon-jets than quark-jets.
\begin{figure*}[h]
\centering
\includegraphics[angle=0,width=0.375\textwidth]{RAA_model_62_line.eps}
\includegraphics[angle=0,width=0.375\textwidth]{RAA_FitResult_line.eps}
\caption{\label{fig:ChHadronSpectraIIresult2}
(color online)
The left panel shows the underlying mono (triangles) and multi (circles)
\raanpart~distributions for PHOBOS Au+Au data at
\snn\,=\,62.4\,GeV~\cite{cite:PHOBOS_RAA_AuAu62}. The solid lines are
meant to guide the eye. The open (closed) symbols represent a pair fit
using centrality bins 0-6\% and 45-50\% (0-6\% and 45-50\%).
The dashed lines depict the underlying mono and multi distributions
from Au+Au collisions at 200\,GeV (PHENIX). In order to better compare
the multi distributions, the green squares represent the 62.4\,GeV data
rescaled to the same \pp~reference (i.e. assumes a 200\,GeV \pp~reference,
not 62.4\,GeV).
The right panel has the same notation as the left, except the data is for
PHOBOS Cu+Cu collision data at 200\,GeV~\cite{cite:PHOBOS_RAA_CuCu}.}
\end{figure*}
It is interesting to note that the disappearance of the away-side
jet in two-particle correlations was observed in a similar
\pT~region (for the associated particles) as the mono
suppression~\cite{cite:STAR_DisappB2BJet}.
In this kinematic region, 1$<$\pT$<$3\,GeV/$c$, we find that
the underlying mono distribution is very small which perhaps hints
that jets are dominantly from the mono (\pp-like) interactions which
occur on the surface. With this assumption, the away-side is then
suppressed (or fully absorbed) within the (multi) medium. At higher
\pT, the mono and multi distributions essentially become the same and
are similar in magnitude to the measured Cronin-enhanced $d$+Au
data~\cite{cite:PHENIX_dAu_Spectra}. It should be noted that in order
to form these spectral distributions, the \ncoll~scale is explicitly
removed. \raanpart~distributions are formed entirely with our
participant-like parameters, thus any large-scale suppression in the
data (using \raa) is actually brought about by scaling
the distributions by \ncoll. In other words, the suppression at
high-\pT~could be entirely an artifact of using the wrong scaling
variable (namely \ncoll). How could \ncoll~be the wrong variable?
Consider a simple picture. In a standard Glauber model calculation,
we assume that each nucleon-nucleon interaction is independent
and that the cross-section for each interaction is the same (i.e.
$\sigma_{NN}$ does not change -- even after 10 collisions).
Once a ``collision'' has occurred one
could surmise that the cross-section is not the same, something has
been lost so the cross-section could be diminished. Further, if one
assumes that the multi component
melts into a single (not discrete) system, then it may be impossible,
in fact improbable, to manifest multiple collisions, or at least
reliably count them. Note that, experimentally, one cannot count
\ncoll, yet one can directly count \npart, or at least derive it
from the data with some certainty.
It has already been observed that by dividing the \snn\,=\,200\,GeV
charged hadron spectra by that from 62.4\,GeV, the ratio (for a given
\pT) is invariant across all centrality bins
measured~\cite{cite:PHOBOS_CuCu_MidRap}. This may already indicate
a simple geometry scaling which factorizes in energy and centrality.
We should therefore expect a similarly successful fit using the mono/multi
approach for 62.4\,GeV
data. Fig.~\ref{fig:ChHadronSpectraII62} shows the same analysis
using PHOBOS charged hadron spectra at \snn\,=\,62.4\,GeV~\cite{cite:PHOBOS_RAA_AuAu62}, where two
centrality bin pairs can be used to predict all other bins.
Fig~\ref{fig:ChHadronSpectraIIresult2} shows the resultant mono
distribution, illustrating the same level of suppression as 200\,GeV,
whereas 62.4\,GeV has a higher multi \raanpart. In the calculation
of \raanpart, the \pp~spectrum is used in the denominator to observe
any differences between A+A and \pp~collisions. To more fairly compare
the two spectra, the green squares represent the same 62.4\,GeV data,
but rescaled to effectively use the \pp~200\,GeV reference. In this
way, we see that the underlying multi spectrum grows with energy.
A similar rescaling of the underlying mono distribution is not performed
as for that case, the relative change to \pp~is important, not the
overall spectrum. It is interesting to note that \raanpart~for
mono-collisions is the same for 200\,GeV and 62.4\,GeV. This
indicates that the level of suppression does not depend on collision
energy (at least for these RHIC energies).
\begin{figure*}[!h]
\centering
\includegraphics[angle=0,width=0.90\textwidth]{makeModel_MonoMultiData2_CuCu.eps}
\caption{\label{fig:ChHadronSpectraIICuCu}
Nuclear modification factor (scaled by \npart~not \ncoll) for
unidentified charged hadrons from PHOBOS (filled band) in Cu+Cu
collisions at \snn\,=\,200\,GeV~\cite{cite:PHOBOS_RAA_CuCu} and
from the model fit using the pair fit method (lines). For the
solid lines, the centrality
bins 0-6\% and 45-50\% were used, dashed lines used centrality
bins 15-25\% and 45-50\%.}
\end{figure*}
\begin{figure*}[!h]
\centering
\includegraphics[angle=0,width=0.90\textwidth]{CuCu_FromAuAuParams.eps}
\caption{\label{fig:ChHadronSpectraIICuCu_AuAuParams}
Nuclear modification factor (scaled by \npart~not \ncoll) for
unidentified charged hadrons from PHOBOS (filled band) in Cu+Cu
collisions at \snn\,=\,200\,GeV using the model fit results for
the mono and multi distributions from the Au+Au analysis.
The dashed (solid) lines use the results from fitting the PHOBOS
(PHENIX) data.}
\end{figure*}
In the multiplicity analysis, we tested the hypothesis that
common underlying distributions for the Au+Au and Cu+Cu data can be used
to represent the data.
It is possible to also perform the same comparisons with the unidentified
charged hadron spectra. In Fig.~\ref{fig:ChHadronSpectraIICuCu}, a pair-fit
using PHOBOS Cu+Cu data at \snn\,=\,200\,GeV~\cite{cite:PHOBOS_RAA_CuCu}
is invoked to extract the underlying mono- and multi-distributions,
which are then used to reconstitute the \raanpart~in Cu+Cu for each centrality
bin. In contrast, Fig.~\ref{fig:ChHadronSpectraIICuCu_AuAuParams}
uses the underlying mono- and multi-distributions derived from Au+Au
collisions to form the \raanpart~for each centrality bin.
In both cases, a reasonable fit is found. For completeness,
the right panel in Fig.~\ref{fig:ChHadronSpectraIIresult2} shows a
comparison between the underlying mono and multi distributions
from Au+Au (symbols) and those derived from Cu+Cu (lines). The
two sets of symbols for Cu+Cu represent the pair fit method using
PHOBOS centrality bins 0-6\% and 45-50\% (open symbols) and 0-6\%
and 45-50\% (closed symbols).
\section{Freeze-out Properties}
As an extension to the low-\pT~particle studies, it has proven
useful to fit the resultant spectra with models to derive freeze-out
properties of the system. Commonly, the kinetic freeze-out
temperature ($T_{\rm kin}$) and radial flow velocity ($\beta$) of the
system are derived from a Blast-wave~\cite{cite:BlastWave} fit to
the low-\pT~spectra of identified pions, kaons, and protons. For a
description see for example
Refs.~\cite{cite:STAR_CuCu_FreezeOut,cite:STAR_AuAu_FreezeOut}.
At kinetic freeze-out all elastic collisions cease and the spectral
shape is fixed. The extracted parameters reflect the state of the
system at that time.
What has been observed, through the Blast-wave fits, is that
$T_{\rm kin}$ and $\beta$ vary quite strongly as a function of
centrality. Here, we address the question whether those results
can be described within the mono/multi framework. In a similar way
to the above analysis, we consider that mono and multi have their
own ``universal'' $T_{\rm kin}$ and $\beta$ values. Using the
least-$\chi^{2}$ method described above, we fit the centrality
dependence of the freeze-out variables and extract those parameters
($Y$ variables in Eqn.\ref{eqn:MonoMulti}) for mono and multi.
The left and right panels of Fig.~\ref{fig:KinFO} shows the result of
fitting the freeze-out temperature ($T_{\rm kin}$) and radial flow velocity
($\beta$), respectively, at \snn\,=\,200 and 62.4\,GeV from STAR~\cite{cite:STAR_AuAu_FreezeOut}. From the fits we find
that the centrality dependence is well described in this framework,
and the extracted $T_{\rm kin}$ and $\beta$ values for the underlying
mono and multi
components are given in Table~\ref{tbl:FreezeOut}. Although there
is some collision energy dependence, the more striking feature is
that the mono results for $T_{\rm kin}$ and $\beta$ are significantly
different from the multi values. A consistent picture emerges
that the multi component has considerably larger $\beta$
values and a freeze-out temperature of about one-half of that of the mono.
Figure~\ref{fig:KinFOCuCu} shows the results of a comparative analysis
using the freeze-out parameters from Cu+Cu collisions at
STAR~\cite{cite:STAR_CuCu_FreezeOut}.
Differences between Au+Au and Cu+Cu evident, especially
at the highest energy. The heavier Au+Au system has a larger $\beta$
and smaller $T_{\rm kin}$ than those extracted from Cu+Cu collisions,
see Fig.~\ref{fig:KinFOCuCu}. Although there are differences, we
did not consider the systematic uncertainties in the data when creating
the underlying distributions for the two collision systems, so it becomes difficult to draw any strong
conclusions. In comparison to the freeze-out
parameters extracted for \pp~collisions, for Au+Au collisions at
200\,GeV, the extracted $T_{\rm kin}$ for the underlying mono component
is about 25\% higher than that from the corresponding \pp~value
(0.127$\pm$0.013~\cite{cite:STAR_AuAu_FreezeOut}). Similarly, the
extracted radial flow velocity is about 35\% lower than the
corresponding \pp~value (0.244$\pm$0.081~\cite{cite:STAR_AuAu_FreezeOut}).
We should note here, that we do not expect the mono to be precisely
as minimum bias \pp. In particular, the extracted underlying
mono distributions for multiplicity and charged hadron spectra are
different from minimum bias \pp, especially in the low-\pT~region.
From this analysis, if we take the Blast-wave model at face value,
it appears as though the kinetic freeze-out of the mono collisions, around the
surface, occurs earlier than the multi, as implied by the different temperatures
at freeze-out. This, in turn, suggests that the kinetic freeze-out
occurs at a later time such that elastic collisions still occur in the
multi after the complete freeze-out of the mono. The radial
velocity, $\beta$, is found to be far greater for the multi, possibly
indicating a more explosive expansion. Thus, one could argue that the
multi system is simply more densely compact. We note that, the
Blast-wave fits are simply reflecting the difference between the
underlying mono and multi spectral shapes.
\begin{figure*}[!bh]
\begin{minipage}{18pc}
\includegraphics[angle=0,width=0.95\textwidth]{KinFOTemp.eps}
\end{minipage}\hspace{2pc}
\begin{minipage}{18pc}
\includegraphics[angle=0,width=0.95\textwidth]{KinFOBeta.eps}
\end{minipage}\hspace{2pc}
\caption{\label{fig:KinFO}
(color online) Least-$\chi^{2}$ fit to the kinetic freeze-out temperature
(left) and radial flow velocity (right) for Au+Au collisions at 200 (filled
band) and 62.4\,GeV (outline band). Dashed (dotted) lines represent the
200\,GeV (62.4\,GeV) fit results. The left (right) pointing arrows depict
the mono (multi) $T_{\rm kin}$ or $\beta$ in each panel.}
\end{figure*}
\begin{figure*}[!bh]
\begin{minipage}{18pc}
\includegraphics[angle=0,width=0.95\textwidth]{KinFOTemp_CuCu.eps}
\end{minipage}\hspace{2pc}
\begin{minipage}{18pc}
\includegraphics[angle=0,width=0.95\textwidth]{KinFOBeta_CuCu.eps}
\end{minipage}\hspace{2pc}
\caption{\label{fig:KinFOCuCu}
(color online) Least-$\chi^{2}$ fit to the kinetic freeze-out temperature
(left) and radial flow velocity (right) for Cu+Cu collisions at 200 (filled
band) and 62.4\,GeV (outline band). Dashed (dotted) lines represent the
200\,GeV (62.4\,GeV) fit results. The left (right) pointing arrows depict
the mono (multi) $T_{\rm kin}$/$\beta$.}
\end{figure*}
In the evolution of the collision, a chemical freeze-out occurs prior
to the kinetic freeze-out. Chemical freeze-out occurs once all inelastic
collisions have ceased; freezing the particle species. It is found that
this temperature ($T_{ch}$) is invariant with
centrality~\cite{cite:STAR_CuCu_FreezeOut}, thus both the mono and
multi must have the same apparent chemical freeze-out temperature.
\begin{table*}[!th]
\caption{\label{tbl:FreezeOut} Mono and multi freeze-out parameters
as derived from the least-$\chi^{2}$ fit to the mid-rapidity data from
STAR.}
\begin{tabular}{| c || c | c || c | c || c | c || c | c |}
\hline
& \multicolumn{4}{c ||}{Au+Au} & \multicolumn{4}{ c |}{Cu+Cu} \\
Energy & \multicolumn{2}{c ||}{mono} & \multicolumn{2}{ c ||}{multi} & \multicolumn{2}{c ||}{mono} & \multicolumn{2}{ c |}{multi} \\
(GeV) & T$_{\rm kin}$ [GeV] & $\beta$ & $T_{\rm kin}$ [GeV] & $\beta$ & $T_{\rm kin}$ [GeV] & $\beta$ & $T_{\rm kin}$ [GeV] & $\beta$ \\\hline
200 & 0.160 & 0.163 & 0.084 & 0.639 & 0.152 & 0.268 & 0.112 & 0.555 \\
62.4 & 0.168 & 0.184 & 0.096 & 0.576 & 0.156 & 0.226 & 0.114 & 0.548 \\ \hline
\end{tabular}
\end{table*}
\section{Summary}
In summary, we have shown that some aspects of heavy ion data can be
reproduced using a simple two-component model, with participant-like
scaling variables. As an initial test of the model, we find that the
multiplicity and unidentified charged hadron spectra can be reproduced
with an underlying gluon-dominated distribution; the latter without
the need for a large suppression at high-\pT. By fitting the data, we
can simultaneously extract underlying distributions for each sub-component
of the collision: the singly-hit mono participants and the multiply-hit
multi. From this, a suppression is observed, however, it is limited to
particles originating at surface-collisions (mono) and mostly in the low-
to intermediate-\pT~region. The component characterizing the multiple
nucleon-nucleon interactions is found to be peaked in the intermediate
\pT~region and becomes flat at the highest measured \pT. With these
two sub-components, we observe that there is a mechanism by which one
can reliably reproduce the multiplicity, spectral shapes, and freeze-out
parameters. These is no need for a large (up to a factor of
five) suppression which is currently used to describe the heavy ion data
using the \ncoll~variable.
At the center of this model picture is a dense (perhaps gluon dominated)
system (multi) which is opaque to the particles produced at the periphery
of the collision (mono). A gluon-dominated picture is not excluded by
other measurements, as discussed, the anomalous increase of
baryons, with respect to mesons, could be expected from a system with a
higher number of gluon-jets than quark-jets, though more details are
outside the scope of the current paper.
To conclude, in our model, we assert that the yield of high-\pT~particles
is not suppressed. The apparent reduction in yields observed at
high-\pT~is in fact due to a different particle production mechanism than
what is expected from minimum bias \pp~interactions. In particular this
mechanism does not produce high-\pT~particles at the expected binary
collision rate, but rather at a slower rate as determined by the
geometry of the collision, in a similar way to that expected from
close-packed gluons.
\begin{acknowledgments}
This work was partially supported by US DOE Grants
DE-FG02-04ER41325,
DE-FG03-86ER40271,
and DE-FG02-94ER40865
\end{acknowledgments}
|
\section{Introduction}
Bismut's derivative formula \cite{Bismut} for diffusion semigroups on Riemannian manifolds, also known as Bismut-Elworthy-Li formula due to \cite{EL}, is a powerful tool for stochastic analysis on Riemannian manifolds. On the other hand, the dimension-free Harnack inequality introduced in \cite{W97} has been efficiently applied to the study of functional inequalities, heat kernel estimates and strong Feller properties
in both finite- and infinite-dimensional models, see \cite{ATW06, ATW09, DRW09, ES, K, LW, Ouyang, ORW, RW, W07, W10, WWX10, WX10, WY10, Z}. These two objects have been well developed in the elliptic setting,
but the study for the degenerate case is far from complete.
It is known that the Bismut type formula can be derived for a class of hypoelliptic diffusion semigroups by using Malliavin calculus (see e.g. \cite[Theorem 10]{AT99}). In this case, since no curvature bound can be used, the derivative formulae are usually less explicit. It is remarkable that in the recent work \cite{Zhang} X. Zhang established an explicit derivative formula for the semigroup associated to degenerate SDEs of type (\ref{1.1}) below (see Section 2 for details). On the other hand, the study of dimension-free Harnack inequality for degenerate diffusion semigroups is very open, except for Ornstein-Uhlenck type semigroups investigated in \cite{ORW}, where the associated stochastic differential equation is linear.
Our strategy is based on coupling, see for example \cite{W10b}, and the main purpose of the paper is thus to construct such a successful coupling using Girsanov transform in the manner of \cite{ATW06} for degenerate diffusion processes, which implies explicit Bismut formula and dimension-free Harnack inequality for degenerate Fokker-Planck equations.
Let us introduce more precisely the framework we will consider. Let $\sigma} \def\ess{\text{\rm{ess}}_t$ be invertible $d\times d$-matrix which is continuous in $t\ge 0$, $A$ be an $m\times d$-matrix with rank $m$, $B_t$ be a $d$-dimensional Brownian motion, and $Z_t\in C^1(\R^m\times\R^d,\R^d)$ which is continuous in $t$. Consider the following degenerate stochastic differential equation on $\R^m\times \R^d$:
\beq\label{1.1} \begin} \def\beq{\begin{equation}} \def\F{\scr F{cases} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= A Y_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t,\\
\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D Y_t= \sigma} \def\ess{\text{\rm{ess}}_t \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_t + Z_t(X_t,Y_t)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t.\end{cases}\end{equation} We shall use $(X_t(x), Y_t(y))$ to denote the solution
with initial data $(x,y)\in \R^m\times\R^d$. For simplicity, we will use $\R^{m+d}$ to stand for $\R^{m}\times \R^d$. Then the solution is a Markov process generated by
$$L_t:=\ff 1 2 \sum_{i,j=1}^d (\sigma} \def\ess{\text{\rm{ess}}_t\sigma} \def\ess{\text{\rm{ess}}^*_t)_{ij}\ff{\pp^2}{\pp y_i\pp y_j} +\sum_{i=1}^d (Z_t(x,y))_j \ff{\pp}{\pp y_j} +\sum_{l=1}^m (Ay)_l \ff{\pp}{\pp x_l}.$$
For any $f\in \B_b(\R^{m+d}),$ the set of all bounded measurable real functions on $\R^{m+d},$ let
$$ P_t f(x,y):= \E f(X_t(x), Y_t(y)),\ \ t\ge 0, (x,y)\in \R^{m+d}.$$ Then $u(t,x,y):=P_t f(x,y)$ solves the degenerate Fokker-Planck type equation
$$\pp_t u(t,x,y)= L_tu(t,\cdot)(x,y).$$
In the case where $m=d$, $\sigma_t=A=I$ and
$$Z_t(x,y)=-\nabla V(x)-cy,$$
this type of equation has recently attracted much interest under the name $``$kinetic Fokker-Planck equation" in PDE, see Villani \cite{Vil09}, or
$``$stochastic damping Hamiltonian system" in probability, see \cite{Wu01,BCG08}, where the long time behavior of $P_t$ has been investigated. In this particular case the invariant probability measure (if it exists) is well known as $\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D x,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D y)=\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-2V(x)-c|y|^2}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D x \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D y$ (up to a constant), and Villani \cite{Vil09} uses this fact to establish hypocoercivity via most importantly an hypoelliptic regularization estimate $H^1\to L^2$. First note that the methodology used there relies heavily on the knowledge of the invariant measure, which we will not need in the present study. Also, his main condition reads as $|\nabla^2V|\le c(1+|\nabla V|)$ preventing exponentially growing potentials, but for parts of our results we do not impose such growing conditions. To allow easier comparison, we will use as running example kinetic Fokker-Planck equation. Let us also mention that we obtain here pointwise estimates, i.e. control of $|\nabla P_t f|$, which allows for example to get uniform bounds when $f$ is initially bounded (exploding when time goes to 0), results that cannot be obtained via Villani's methodology.
In the following three sections, we will investigate pointwise regularity estimates by establishing derivative formula, gradient estimate and Harnack inequality for $P_t$.
\section{Derivative formulae}
Since $A$ has rank $m$, we have $d\ge m$ and for any $h_1\in \R^m$, the set
$$A^{-1} h_1:=\{z\in\R^d:\ Az =h_1\}\ne\emptyset.$$
For any $h_1\in\R^m$, let
$$|A^{-1}h_1|=\inf\{|z|:\ z\in A^{-1}h_1\}.$$ Then it is clear that
$$\|A^{-1}\|:=\sup\big\{|A^{-1}h_1|:\ h_1\in \R^m, |h_1|\le 1\big\}<\infty.$$
We shall use $|\cdot|$ to denote the absolute value and the norm in Euclidean spaces, and use $\|\cdot\|$ to denote
the operator norm of a matrix. For $h\in \R^{m+d}$, we use $D_h$ to stand for the directional derivative along $h$.
Before move on, let us first mention the Bismut formula derived in \cite{Zhang}. We call a $C^2$-function $W$ on $\R^{m+d}$ a Lyapunov function, if $W\ge 1$ having compact level sets. The following result is reorganized from \cite[Theorem 3.3]{Zhang}. For $h\in \R^{m+d}$, let $\nabla} \def\pp{\partial} \def\EE{\scr E_h$ denote the directional derivative along $h$.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}[\cite{Zhang}] Let $t>0,$ $m=d$ and $A=I$. Assume that there exist a Lyapunov function $W$ and some constants $C>0, \aa\in [0,1], \ll\ge 0$
such that for $s\in [0,t]$
\beq\label{Z0} L_s W\le C W, \ \ |\nabla} \def\pp{\partial} \def\EE{\scr E W|^2\le CW^{2-\aa} \end{equation} and
\beq\label{Z0'} \begin} \def\beq{\begin{equation}} \def\F{\scr F{cases}|\nabla} \def\pp{\partial} \def\EE{\scr E Z_s|\le C W^\ll, \\
\<y-\tilde} \def\Ric{\text{\rm{Ric}} y,Z_s(x,y)-Z_t(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\>\le C|(x-\tilde} \def\Ric{\text{\rm{Ric}} x, y-\tilde} \def\Ric{\text{\rm{Ric}} y)|^2 \big\{W(x,y)^\aa+W(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)^\aa\big\} \end{cases}\end{equation} hold for $(x,y),(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\in\R^{m+d}$. Then for any $h=(h_1,h_2)\in \R^{m+d}$ and $f\in \B_b(\R^{m+d})$,
$$\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f= \ff 1 t \E\bigg\{ f(X_t,Y_t)\int_0^t\Big\<\sigma} \def\ess{\text{\rm{ess}}^{-1}_s\big\{\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta_s} Z_s(X_s,Y_s)-\gg_1'(s)h_1+\gg_2'(s)h_2\big\}, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\Big\>\bigg\}$$ holds, where
$$\gg_1(s)= 2(t-2s)^++s-t, \ \ \gg_2(s)= \ff 4 t \{s\land (t-s) \}$$ and
$$\Theta_s= \bigg(h_1\int_0^s\gg_1(r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r + h_2 t+ h_2 \int_0^s \gg_2(r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r,\ \gg_1(s)h_1-\gg_2(s)h_2\bigg).$$\end{thm}
In particular, this result applies to $W(x,y)=1+|x|^2+|y|^2$ and $\aa=0$ provided $|\nabla} \def\pp{\partial} \def\EE{\scr E Z|$ is bounded. In general, however, the assumption $|\nabla} \def\pp{\partial} \def\EE{\scr E W|^2\le C W^{2-\aa}$ excludes exponential choices of $W$ like $\exp[|x|^l+|y|^m]$ for $l\lor m>1$, which is exactly the correct Lyapunov function in the study of kinetic Kokker-Planck equation (see
Example 2.1 below). In this section, we aim to present a more general version of the derivative formula without this condition.
Let us introduce now the assumption that we will use in the sequel:
\paragraph{(A)} \emph{There exists a constant $C>0$ such that $L_s W\le CW$ and
$$|Z_s(\x)-Z_s(\y)|^2\le C|\x-\y|^2 W(\y),\ \ \x,\y\in\R^{m+d}, |\x-\y|\le 1$$
hold for some Lyapunov function $W$ and $s\in [0,t]$.}
\
Note that condition $L_s W\le CW$, included also in (\ref{Z0}), is normally a easy to check condition in applications. Although the second condition in {\bf (A)} might be stronger than (\ref{Z0'}), it is a natural condition to exchange the order of the expectation and the derivative by using the dominated convergence theorem, which is however missed in \cite{Zhang} (see line 4 on page 1942 therein). Most importantly, the second condition in (\ref{Z0}) is now dropped, so that we are able to treat highly non-linear drift $Z$ as in Examples 2.1 and 4.1 below.
The main result in this section provides various different versions of derivative formula by making different choices of the pair functions $(u,v)$.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}\label{T1.1} Assume {\bf (A)}. Then the process $(X_t,Y_t)_{t\ge0}$ is non-explosive for any initial point in $\R^{m+d}$. Moreover, let $t>0$ and $u,v\in C^2([0,t])$ be such that
\beq\label{1.0} u(t)=v'(0)=1,\ \ \ u(0)=v(0)=u'(0)=u'(t)=v'(t)=v(t)=0.\end{equation} Then for any $h=(h_1,h_2)\in \R^m\times \R^d$ and $z\in A^{-1}h_1:=\{z\in \R^d:\ Az=h_1\},$
\beq\label{1.2} \nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f = \E\bigg\{f(X_t, Y_t)
\int_0^t \Big\<\sigma} \def\ess{\text{\rm{ess}}_s^{-1}\big\{u''(s) z -v''(s)h_2
+(\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z_s)(X_s,Y_s)\big\},\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\Big\>\bigg\} \end{equation} holds for $f\in \B_b(\R^{m+d})$, where
$$\Theta(h,z,s)= \big(\{1-u(s)\}h_1 +v(s)Ah_2,\ v'(s)h_2-u'(s)z\big).$$
\end{thm}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} The non-explosion follows since $L_sW\le CW$ implies
\beq\label{EXP00} \E W(X_s,Y_s)\le W\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{Cs},\ \ s\in [0,t], (x,y)\in \R^{m+d}.\end{equation}
To prove (\ref{1.2}), we make use of the coupling method with control developed in \cite{ATW06}. Since the process is now degenerate, the construction of coupling is highly technical: we have to force the coupling to be successful before a fixed time by using a lower dimensional noise.
Let $t>0,(x,y),h=(h_1,h_2)\in \R^{m+d}$ and $z\in A^{-1}h_1$ be fixed. Simply denote $(X_s,Y_s)=(X_s(x), Y_s(y))$. From now on, let
$$\vv_0= \inf_{s\in [0,t]}\ff 1 {1\lor |\Theta(h,z,s)|}>0,$$ so that $\vv_0 |\Theta(h,z,s)|\le 1$ for $s\in [0,t]$. For any $\vv\in (0,\vv_0),$
let $(X_s^\vv, Y_s^\vv)$ solve the equation
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation} \label{CC}\begin} \def\beq{\begin{equation}} \def\F{\scr F{cases} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_s^\vv= A Y_s^\vv\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s,\ \ X_0^\vv= x+\vv h_1,\\
\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D Y_s^\vv =\sigma} \def\ess{\text{\rm{ess}}_s\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s +Z_s(X_s,Y_s)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s+\vv\{v''(s)h_2-u''(s)z\}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s,\ \ Y_0^\vv=y+\vv h_2.\end{cases}\end{equation} By (\ref{1.0}) and noting that $Az=h_1$, we have
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{2.1}\begin} \def\beq{\begin{equation}} \def\F{\scr F{cases} Y_s^\vv=Y_s +\vv v'(s)h_2-\vv u'(s)z,\\
X_s^\vv= x+\vv h_1 + A\int_0^s Y_r^\vv\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r= X_s +\vv \{1-u(s)\}h_1 +\vv v(s)Ah_2.\end{cases}\end{equation}
Due to (\ref{1.0}), this in particular implies
\beq\label{2.2} (X_t,Y_t)=( X_t^\vv, Y_t^\vv),\end{equation}
and also that
\beq\label{2.2'}( X_s^\vv, Y_s^\vv)=(X_s,Y_s)+\vv\Theta(h,z,s),\ \ s\in [0,t].\end{equation}
On the other hand, let
$$\xi^\vv_s= Z(X_s,Y_s)-Z( X_s^\vv, Y_s^\vv) +\vv v''(s) h_2 -\vv u''(s)z,\ \ s\in [0,t]$$ and
\beq\label{WF0}R_s^\vv= \exp\bigg[-\int_0^{s}\<\sigma} \def\ess{\text{\rm{ess}}^{-1}_s\xi_r^\vv, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_r\> -\ff 1 2 \int_0^{s}|\sigma} \def\ess{\text{\rm{ess}}^{-1}_s\xi_r^\vv|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\bigg], \ \ s\in [0,t].\end{equation} We have
$$\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D Y_s^\vv =\sigma} \def\ess{\text{\rm{ess}}_s \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s^\vv +Z_s( X_s^\vv, Y_s^\vv)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$$ for
$$ B_s^\vv:= B_s+ \int_0^s\sigma} \def\ess{\text{\rm{ess}}^{-1}_s\xi^\vv_r\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r,\ \ s\in [0,t],$$ which is $d$-dimensional Brownian motion under the probability measure
$\Q_\vv:= R^\vv_t\P$ according to Lemma \ref{L1} below and the Girsanov theorem. Thus, due to (\ref{2.2}) we have
$$ P_t f((x, y)+\vv h) =\E_{\Q_\vv} f( X_t^\vv, Y_t^\vv) =\E [R^\vv_t f(X_t,Y_t)].$$ Since $P_tf(x,y)=\E f(X_t,Y_t),$ we arrive at
$$ P_tf((x,y)+ \vv h) -P_t f(x,y)= \E[(R^\vv_t-1)f(X_t, Y_t)].$$ The proof is then completed by Lemma \ref{L2}.
\end{proof}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L1} If {\bf (A)} holds, then
$$\sup_{s\in [0,t],\vv\in (0,\vv_0)} \E \big(R_s^\vv\log R_s^\vv\big)<\infty.$$Consequently, for each $\vv\in (0,\vv_0)$,
$(R_s^\vv)_{s\in [0,t]}$ is a uniformly integrable martingale. \end{lem}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}
$$\tau_n= \inf\{t\ge 0: |X_t(x)|+|Y_t(y)|\ge n\},\ \ n\ge 1.$$ Then $\tau_n\uparrow\infty$ as $n\uparrow\infty.$
By the Girsanov theorem, $(R_{s\land\tau_n})_{s\in [0,t]}$ is a martingale and $\{B_s^\vv: 0\le s\le t\land \tau_n\}$ is a Brownian motion under the probability measure $\Q_{\vv,n}:=R_{t\land\tau_n}^\vv\P$. Noting that
$$\log R_{s\land \tau_n}^\vv =- \int_0^{s\land\tau_n}\<\sigma} \def\ess{\text{\rm{ess}}^{-1}_r\xi_r^\vv, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_r^\vv\> +\ff 1 2 \int_0^{s\land\tau_n}|\sigma} \def\ess{\text{\rm{ess}}^{-1}_r\xi_r^\vv|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r, \ \ s\in [0,t],$$
where the stochastic integral is a $\Q_{\vv,n}$-martingale, we have
\beq\label{NNm}
\E[R_{s\land\tau_n}^\vv\log R_{s\land\tau_n}^\vv]=\E_{\Q_{\vv,n}}[\log R_{s\land\tau_n}^\vv]\le
\ff 1 2\E_{\Q_{\vv,n}} \int_0^{t\land\tau_n}|\sigma} \def\ess{\text{\rm{ess}}^{-1}_r\xi_r^\vv|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r,\ \ s\in [0,t].\end{equation}
Noting that by {\bf (A)} and (\ref{2.2'})
\beq\label{NNm2} |\sigma} \def\ess{\text{\rm{ess}}^{-1}_r \xi_r^\vv|^2\le c \vv^2 W(X_r^\vv, Y_r^\vv),\ \ r\in [0,t]\end{equation} holds for some constant $c>0$, and moreover under the probability measure $\Q_{\vv,n}$
the process $(X_s^\vv, Y_s^\vv)_{s\le t\land\tau_n}$ is generated by $L_s$, $L_sW\le CW$ implies
\beq\label{EXP0} \E_{\Q_{\vv,n} } \int_0^{s\land\tau_n}W(X_r^\vv,Y_r^\vv)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \le \int_0^s \E_{\Q_{\vv} } W(X_r^\vv,Y_r^\vv)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\le W(X_0^\vv,Y_0^\vv)\int_0^t\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{Cr}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r. \end{equation}
Combining this with (\ref{NNm}) we obtain
\beq\label{WF*}\E[R_{s\land\tau_n}^\vv\log R_{s\land\tau_n}^\vv]\le c,\ \ s\in [0,t], \vv\in (0,\vv_0),n\ge 1\end{equation} for some constant $c>0.$
Since for each $n$ the process $(R_{s\land\tau_n}^\vv)_{ s\in [0,t]}$ is a martingale, letting $n\to\infty$ in the above inequality we complete the proof.\end{proof}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{lem}\label{L2} If {\bf (A)} holds then the family
$\big\{\ff{|R_t^\vv -1|}\vv \big\}_{\vv\in (0,\vv_0)}$ is uniformly integrable w.r.t. $\P$. Consequently,
\beq\label{FF} \lim_{\vv\to 0} \ff{R_t^\vv-1} \vv = \int_0^t \Big\<\sigma} \def\ess{\text{\rm{ess}}_s^{-1}\big\{u''(s) z -v''(s)h_2 +(\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z)(X_s(x),Y_s(y))\big\},\, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\Big\> \end{equation} holds
in $L^1(\P).$
\end{lem}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Let $\tau_n$ be in the proof of Lemma \ref{L1} and let
$$N_s^\vv= \sigma} \def\ess{\text{\rm{ess}}_s^{-1} \big\{\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z_s(X_s^\vv,Y_s^\vv)+u''(s)z-v''(s)h_2\big\},\ \ s\in [0,t], \vv\in (0,\vv_0).$$ By {\bf (A)} and (\ref{NNm2}), there exists a constant $c>0$ such that
\beq\label{WFF} \big|\< N_s^\vv, \sigma} \def\ess{\text{\rm{ess}}_s^{-1} \xi_s^\vv\>\big|\le \vv |N_s^\vv|^2+\vv^{-1}|\sigma} \def\ess{\text{\rm{ess}}^{-1}_s\xi_s^\vv|^2\le c\vv W(X_s^\vv,Y_s^\vv),\ \
\vv\in (0,\vv_0), s\in [0,t].\end{equation}
Since $\nabla} \def\pp{\partial} \def\EE{\scr E Z$ is locally bounded, it follows from (\ref{2.2'}) and (\ref{WF0}) that
$$ \ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D}{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\vv} R_{t\land \tau_n}^\vv = R_{t\land\tau_n}^\vv\bigg\{\int_0^{t\land\tau_n}\<N_s^\vv, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\> + \int_0^{t\land \tau_n}\<N_s^\vv, \sigma} \def\ess{\text{\rm{ess}}_s^{-1} \xi_s^\vv\>\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg\},\ \ \vv\in (0,\vv_0), n\ge 1.$$
Combining this with (\ref{WFF}) we obtain
$$ \ff{|R_{t\land\tau_n}^\vv -1|}\vv \le \ff 1 \vv \int_0^\vv R_{t\land\tau_n}^r\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \int_0^{t\land\tau_n}\<N_s^r, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\>
+ c\int_0^{\vv_0} R_{t\land \tau_n}^r\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \int_0^{t\land\tau_n} W(X_s^r,Y_s^r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$$ for $\vv\in (0,\vv_0),n\ge 1.$
Noting that
under $\Q_{r}$ the process $(X_s^r,Y_s^r)_{s\in [0,t]}$ is generated by $L_s$, by (\ref{EXP00}) we have
$$\E \int_0^{\vv_0} R_{t}^r\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \int_0^{t}W(X_s^r,Y_s^r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s= \int_0^{\vv_0} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r \int_0^t \E_{\Q_r} W(X_s^r,Y_s^r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s <\infty.$$ Thus, for the first assertion it remains to show that the family
$$\eta_{\vv,n}:= \ff 1 \vv \int_0^\vv R_{t\land \tau_n}^r |\Xi_{t,n}|(r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r,\ \ \vv\in (0,\vv_0), n\ge 1$$ is uniformly integrable, where
$$\Xi_{t,n}(r):= \int_0^{t\land\tau_n}\<N_s^r, \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\>.$$
Since $r\log^{1/2} (\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega+r)$ is increasing and convex in $r\ge 0$, by the Jensen inequality,
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\E\big\{\eta_{\vv,n} \log^{1/2}(\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega+\eta_{\vv,n})\big\} \\
&\le \ff 1\vv \int_0^\vv \E\Big\{R_{t\land\tau_n}^r |\Xi_{t,n}|(r)\log^{1/2}\big(\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega + R_{t\land\tau_n}^r |\Xi_{t,n}|(r)\big)\Big\}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\\
&\le \ff 1 \vv \int_0^\vv \E\Big\{R_{t\land\tau_n}^r |\Xi_{t,n}|(r)^2 + R_{t\land\tau_n}^r \log\big(\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega + R_{t\land\tau_n}^r |\Xi_{t,n}|(r)\big)\Big\}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\\
&\le \ff 1\vv\int_0^\vv \E\Big\{c+ 2 R_{t\land\tau_n}^r |\Xi_{t,n}|(r)^2 + R_{t\land\tau_n}^r\log R_{t\land\tau_n}^r\Big\}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\end{split}\end{equation*} holds for some constant $c>0$. Combining this with (\ref{WF*}) and noting that (\ref{WFF}) and (\ref{EXP0}) imply
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\E \big\{R_{t\land\tau_n}^r |\Xi_{t,n}|(r)^2\big\}= \E_{\Q_{r,n}}\bigg( \int_0^{t\land\tau_n}
\big\<N_s^r,\ \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s^r\big\>\bigg)^2
=\E_{\Q_{r,n}}\int_0^{t\land\tau_n}|N_s^r|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \\
&\le c \E_{\Q_{r,n}}\int_0^{t\land \tau_n} W(X_s^r, Y_s^r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\le c',\ \ \ n\ge 1, r\in(0,\vv_0) \end{split}\end{equation*} for some constants $c,c'>0$, we conclude that $\{\eta_{\vv,n}\}_{\vv\in (0,\vv_0),n\ge 1}$ is uniformly integrable. Thus, the proof of the first assertion is finished.
Next, by {\bf (A)} and (\ref{2.2'}) we have
$$\lim_{\vv\to 0} \Big|\ff {\xi_s^\vv}\vv + (\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z)(X_s,Y_s)+u''(s)z-v''(s)h_2\Big|=0.$$ Moreover, for each $n\ge 1$ this sequence
is bounded on $\{\tau_n\ge t\}$. Thus, $(\ref{FF})$ holds a.s. on $\{\tau_n\ge t\}$. Since $\tau_n\uparrow\infty$, we conclude that (\ref{FF}) holds a.s.
Therefore, it also holds on $L^1(\P)$ since $\{\ff{R_t^\vv-1}\vv\}_{\vv\in (0,1)}$ is uniformly integrable according to the first assertion. \end{proof}
To conclude this section, we present an example of kinetic Fokker-Planck equation for which $W$ is an exponential function so that (\ref{Z0'}) fails true but {\bf (A)} is satisfied.
\paragraph{Example 2.1}(Kinetic Fokker-Planck equation)
Let $m=d $ and consider
\beq\label{kfp} \begin} \def\beq{\begin{equation}} \def\F{\scr F{cases} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= Y_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t,\\
\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D Y_t= \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_t -\nabla V(X_t)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t-Y_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t\end{cases}\end{equation}
for some $C^2$-function $V\ge 0$ with compact let sets. Let $W(x,y)=\exp[2V(x)+|y|^2]$. We easily get that
$L W= d W.$ Thus, it is easy to see that {\bf (A)} holds for e.g. $V(x)= (1+|x|^2)^l$ or even $V(x)=\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{(1+|x|^2)^l}$ for some constant $l\ge 0$. Therefore, by Theorem \ref{T1.1} the derivative formula (\ref{1.2}) holds for $(u,v)$ satisfying (\ref{1.0}).\\
Note that Villani \cite[th. A.8]{Vil09} has a crucial assumption: $|\nabla^2 V|\le C(1+|\nabla V|)$ which prevents potential behaving as $V(x)=\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{(1+|x|^2)^l}$. Note also that the previous arguments do not rely on the explicit knowledge of an invariant probability measure, which is crucial in Villani's argument.
\section{Gradient estimates}
In this section we aim to derive gradient estimates from the derivative formula (\ref{1.2}). For simplicity, we only consider the time-homogenous case that $\sigma} \def\ess{\text{\rm{ess}}$ and $Z$ are independent of $t$. In general, we have the following result.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{prp} \label{P3.1} Assume {\bf (A)} and let $(u,v)$ satisfy $(\ref{1.0})$. Then for any $f\in \B_b(\R^{m+d}), t>0$ and $h=(h_1,h_2)\in\R^{m+d}$,
$z\in A^{-1} h_1$,
\beq\label{3.1} |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|^2 \le \|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2(P_t f^2) \E \int_0^t \big|u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{equation}If $f\ge 0$ then
for any $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho>0$,
\beq\label{3.2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split}& |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\big\{P_t (f\log f)-(P_t f)\log P_t f\big\}\\
&\qquad+\ff{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_t f}2 \log \E\exp\bigg[\ff {2\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2} {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2} \int_0^t\big|u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg].\end{split}\end{equation}
\end{prp}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Let $M_t= \int_0^t \big\<\sigma} \def\ess{\text{\rm{ess}}^{-1} \big\{u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big\},\ \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\big\>.$ By
(\ref{1.2}) and the Schwartz inequality we obtain
$$ |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|^2 \le (P_t f^2) \E M_t^2 \le \|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2(P_t f^2) \E \int_0^t \big|u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.$$ That is, (\ref{3.1}) holds. Similarly, (\ref{3.2}) follows from (\ref{1.2}) and the Young inequality (cf. \cite[Lemma 2.4]{ATW09}):
$$|\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f| \le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t (f\log f)-(P_t f)\log P_t f\big\}+\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \log\E\exp\Big[\ff{M_t}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\Big]$$ since
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\E\exp\Big[\ff{M_t}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\Big]\le \bigg(\E\exp\Big[\ff{2\<M\>_t}{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2}\Big]\bigg)^{1/2}\\
&\le \bigg(\E\exp\bigg[\ff {2\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2} {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2} \int_0^t\big|u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\bigg)^{1/2}.\end{split}\end{equation*}\end{proof}
To derive explicit estimates, we will take the following explicit choice of the
pair $(u, v)$:
\beq\label{uv1}u(s)= \ff{s^2(3t-2s)}{t^3},\ \ \ v(s)= \ff{s(t-s)^2}{t^2},\ \ \ \ s\in [0,t],\end{equation} which satisfies (\ref{1.0}). In this case
we have
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{uv2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &u'(s)=\ff{6s(t-s)}{t^3},\ u''(s)= \ff{6(t-2s)}{t^3}, \ v'(s)=\ff{(t-s)(t-3s)}{t^2},\\
&\ v''(s)=\ff{2(3s-2t)}{t^2},\ 1-u(s)= \ff{(t-s)^2(t+2s)}{t^3} ,\ \ s\in [0,t].\end{split}\end{equation}
In this case, Proposition \ref{P3.1} holds for
\beq\label{uv}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &u''(s)z-v''(s)h_2=\LL(h,z,s):= \ff{6(t-2s)}{t^3}z +\ff{2(2t-3s)}{t^2}h_2,\\
&\Theta(h,z,s)= \\
&\ \Big(\ff{(t-s)^2(t+2s)}{t^3}h_1+\ff{s(t-s)^2}{t^2} A h_2, \ff{(t-s)(t-3s)}{t^2}h_2-\ff{6s(t-s)}{t^3}z\Big).\end{split}\end{equation}
Below we consider the following three cases respectively:
\begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate}\item[(i)] $|\nabla} \def\pp{\partial} \def\EE{\scr E Z|$ is bounded;
\item[(ii)] $|\nabla} \def\pp{\partial} \def\EE{\scr E Z|$ has polynomial growth and $\<Z(x,y), y\>\le C(1+|x|^2+|y|^2)$ holds for some constant $c>0;$
\item[(iii)] A more general case including the kinetic Fokker-Planck equation.\end{enumerate}
\subsection{Case (i): $|\nabla} \def\pp{\partial} \def\EE{\scr E Z|$ is bounded}
In this case {\bf (A)} holds for e.g. $W(x,y)=1+|x|^2+|y|^2$, so that Proposition \ref{P3.1} holds for $u''(s)z-v''(s)h_2$ and $\Theta(h,z,s)$ given in (\ref{uv}). From this specific choice of $\Theta(h,z,s)$
we see that $\nabla} \def\pp{\partial} \def\EE{\scr E^x Z$ and $\nabla} \def\pp{\partial} \def\EE{\scr E^y Z$ will lead to different time behaviors of $\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f$. So, we adopt the condition
\beq\label{NZ} |\nabla} \def\pp{\partial} \def\EE{\scr E^x Z(x,y)|\le K_1,\ \ |\nabla} \def\pp{\partial} \def\EE{\scr E^y Z(x,y)|\le K_2,\ \ (x,y)\in \R^{m+d}\end{equation} for some constants $K_1,K_2\ge 0,$ where $\nabla} \def\pp{\partial} \def\EE{\scr E^x$ and $\nabla} \def\pp{\partial} \def\EE{\scr E^y$ are the gradient operators w.r.t. $x\in\R^m$ and $y\in\R^d$ respectively.
Moreover, for $t>0$ and $r_1,r_2\ge 0$, let
$$\Psi_t(r_1,r_2)=\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2t\bigg\{r_1\Big(\ff {6\|A^{-1}\|}{t^2} +K_1 +\ff{3K_2\|A^{-1}\|}{2t}\Big)
+r_2 \Big(\ff 4 t +\ff{4K_1 t\|A\|}{27}+K_2\Big)\bigg\}^2$$ and
\beq\label{Phi} \Phi_t(r_1,r_2)= \inf_{s\in (0,t]} \Psi_s(r_1,r_2).\end{equation}
In the following result the inequality (\ref{1.3}) corresponds to the pointwise estimate of the $H^1\to L^2$ regularization investigated in Villani \cite[Th. A.8]{Vil09}, while (\ref{1.5}) corresponds to the pointwise estimate of the regularization $``$Fisher information to entropy" \cite[Th A.18]{Vil09}.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor} \label{C2} Let $(\ref{NZ})$ hold for some constants $K_1,K_2\ge 0$. Then for any $t>0, h=(h_1,h_2)\in\R^{m+d}$,
\beq\label{1.3} |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|^2 \le (P_t f^2)\Phi_t(h_1,h_2),\ \ f\in \B_b(\R^{m+d}).
\end{equation} If $f\ge 0$, then
\beq\label{1.4}|\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t (f\log f)- (P_t f)\log (P_tf)\big\}
+\ff {P_tf}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \Phi_t(h_1,h_2)\end{equation} holds for all $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho>0,$ and consequently
\beq\label{1.5}|\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|^2\le 4\Phi_t(h_1,h_2) \big\{P_t (f\log f)- (P_t f)\log (P_tf)\big\} P_t f. \end{equation}
\end{cor}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof}Let $z$ be such that $|z|=|A^{-1}h_1|\le \|A^{-1}\||h_1|,$ and take
\beq\label{Eta} \eta_s= \LL(h,z,s) +\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s(x),Y_s(y)).\end{equation}
By (\ref{3.1}),
\beq\label{3.1'} |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_tf (x,y)|^2\le \|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2(P_tf^2)(x,y)\E\int_0^t|\eta_s|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{equation}
Since (\ref{NZ}) implies $|\nabla} \def\pp{\partial} \def\EE{\scr E_hZ|\le K_1|h_1|+K_2|h_2|$, it follows that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\eta_s|&\le \Big|\ff{6(t-2s)}{t^3}z +\ff{2(2t-3s)}{t^2}h_2\Big| +K_1\Big|\ff{(t-s)^2(t+2s)}{t^3}h_1+\ff{s(t-s)^2}{t^2} A h_2\Big|\\
&\qquad\qquad
+K_2\Big|\ff{(t-s)(t-3s)}{t^2}h_2-\ff{6s(t-s)}{t^3}z\Big|\\
&\le |h_1|\Big(\ff {6\|A^{-1}\|}{t^2} +K_1 +\ff{3K_2\|A^{-1}\|}{2t}\Big)
+|h_2| \Big(\ff 4 t +\ff{4K_1 t\|A\|}{27}+K_2\Big).\end{split}\end{equation*} Then
\beq\label{3.2'}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} \int_0^t | \eta_s|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\le&
t \bigg\{|h_1|\Big(\ff {6\|A^{-1}\|}{t^2}+K_1 +\ff{3K_2\|A^{-1}\|}{2t}\Big)\\&\qquad\qquad\qquad
+|h_2| \Big(\ff 4 t +\ff{4K_1 t\|A\|}{27}+K_2\Big)\bigg\}^2.\end{split}\end{equation}
Combining this with (\ref{3.1'}) we obtain
$$ |\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|^2 \le (P_t f^2)\Psi_t(|h_1|,|h_2|).$$
Therefore, for any $s\in (0,t]$ by the semigroup property and the Jensen inequality one has
$$|\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|^2=|\nabla} \def\pp{\partial} \def\EE{\scr E P_s(P_{t-s} f)|^2\le \Psi_s(|h_1|,|h_2|)P_s(P_{t-s}f)^2\le \Psi_s(|h_1|,|h_2|)P_tf^2.$$
This proves (\ref{1.3}) according to (\ref{Phi}).
To prove (\ref{1.4}) we let $f\ge 0$ be bounded. By (\ref{3.2}),
\beq\label{EG}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\nabla} \def\pp{\partial} \def\EE{\scr E_hP_tf|\le &\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t(f\log f)-(P_tf)\log (P_tf)\big\}\\
&+\ff {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_tf} 2 \log \E\,\exp\bigg[\ff {2\|\sigma} \def\ess{\text{\rm{ess}}^{-1}\|^2} {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2} \int_0^t |\eta_s|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg].\end{split}\end{equation}
Combining this with (\ref{3.2'}) we obtain
$$ |\nabla} \def\pp{\partial} \def\EE{\scr E_hP_tf|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t(f\log f)-(P_tf)\log (P_tf)\big\}
+\ff {P_tf}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \Psi_t(|h_1|,|h_2|).$$ As observed above, by the semigroup property and the Jensen inequality, this implies (\ref{1.4}).
Finally, minimizing the right hand side of (\ref{1.4}) in $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho>0$, we obtain
$$|\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|\le 2 \ss{\Phi_t (|h_1|,|h_2|)\{P_t(f\log f)-(P_t f)\log P_t f\}P_t f}.$$ This is equivalent to (\ref{1.5}). \end{proof}
\subsection{Case (ii)}
Assume there exists $l> 0$ such that
\paragraph{(H)} (i) $\<Z(x,y), y\> \le C(|x|^2+|y|^2+1),\ (x,y)\in \R^{m+d}$;
\ \ \ (ii) $|\nabla} \def\pp{\partial} \def\EE{\scr E Z|(x,y):=\sup\{|\nabla} \def\pp{\partial} \def\EE{\scr E_h Z|(x,y):\ |h|\le 1\}\le C(1+|x|^2+|y|^2)^l,\ (x,y)\in\R^{m+d}.$
\
It is easy to see that {\bf (H)} implies {\bf (A)} for $W(x,y)=(1+|x|^2+|y|^2)^{2l},$ so that Proposition \ref{P3.1} holds for $u''(s)z-v''(s)h_2$ and $\Theta(h,z,s)$ given in (\ref{uv}).
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor} \label{C3} Let {\bf (H)} hold. \begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate} \item[$(1)$] There exists a constant $c>0$ such that
$$ |\nabla} \def\pp{\partial} \def\EE{\scr E P_tf|^2 (x,y)\le \ff{c}{(t\land 1)^3} P_t f^2(x,y),\ \ f\in \B_b(\R^{m+d}),\ \ t>0, (x,y)\in \R^{m+d}.$$
\item[$(2)$] If $l<\ff 1 2$, then there exists a constant $c>0$ such that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|(x,y)\le &\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t (f\log f)- (P_t f)\log (P_tf)\big\}(x,y)\\
&+ \ff {c P_t f(x,y)} {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho (t\land 1)^4}\big\{(|x|^2+|y|^2)^{2l} + (\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho(1\land t)^2)^{4(l-1)/(1-2l)}\big\}\end{split}\end{equation*} holds for all $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho>0$ and positive $f\in \B_b(\R^{m+d})$ and $(x,y)\in\R^{m+d}$.
\item[$(3)$] If $l=\ff 1 2$, then there exist two constants $c,c'>0$ such that for any $t>0$ and $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\ge t^{-2}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{c(1+t)}$,
$$ |\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|(x,y)\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\big\{P_t(f\log f)-(P_t f)\log P_t f\big\}(x,y) + \ff{c' P_tf(x,y)}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big( 1+|x|^2+|y|^2\big)$$ holds for all positive $f\in \B_b(\R^{m+d}) $ and $(x,y)\in \R^{m+d}.$
\end{enumerate}
\end{cor}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} As observed in the proof of Corollary \ref{C2}, we only have to prove the results for $t\in (0,1].$
(1) It is easy to see that $\eta_s$ in the proof of Corollary \ref{C2} satisfies
\beq\label{si}|\sigma} \def\ess{\text{\rm{ess}}^{-1}\eta_s|^2\le c_1(t^2+t^{-4}) |h|^2 (1+|X_s(x)|^2+|Y_s(y)|^2)^{2l}\end{equation}
for some constant $c_1>0$. Thus, the first assertion follows from (\ref{3.1'}) and Lemma \ref{L1}.
(2) Let {\bf (H)} hold for some $l\in (0,1/2)$. Then $$L(1+|x|^2+|y|^2)^{2l}\le c_2 (1+|x|^2+|y|^2)^{2l}$$ holds for some constant $c_2>0$.
Let $(X_s, Y_s)= (X_s(x), Y_s(y))$. By the It\^o formula, we have
$$\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D (1+|X_s|^2+|Y_s|^2)^{2l} \le 4l (1+|X_s|^2+|Y_s|^2)^{2l-1}\<Y_s,\sigma} \def\ess{\text{\rm{ess}}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\> +c_2 (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.$$ Thus,
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\big\{\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\big\}\\
& \le 4l \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l-1}\<Y_s,\sigma} \def\ess{\text{\rm{ess}}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\> -
\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{split}\end{equation*} Therefore, for any $\ll>0$,
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{EXP}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s}\\
& \le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll (1+|x|^2+|y|^2)^{2l}}
\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{4\ll l\int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l-1}\<Y_s,\sigma} \def\ess{\text{\rm{ess}} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_s\>}\\
&\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll (1+|x|^2+|y|^2)^{2l}}\Big\{\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{16\ll^2l^2\|\sigma} \def\ess{\text{\rm{ess}}\|^2 \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-2(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2(2l-1)}|Y_s|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s}\Big\}^{1/2}\\
&\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll (1+|x|^2+|y|^2)^{2l}}\Big\{\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{16\ll^2l^2\|\sigma} \def\ess{\text{\rm{ess}}\|^2 \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{4l-1}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s}\Big\}^{1/2}.\end{split}\end{equation}
On the other hand, since $l<\ff 1 2$ implies $4l-1< 2l$, there exists a constant $c_3>0$ such that
$$16\ll^2l^2\|\sigma} \def\ess{\text{\rm{ess}}\|^2r^{4l-1}\le \ll r^{2l}+ c_3 \ll^{(3-4l)/(1-2l)},\ \ \ r\ge 0.$$ Combining this with (\ref{EXP}) we arrive at
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & \E \exp\bigg[\ll \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\\
& \le \exp\Big[\ll (1+|x|^2+|y|^2)^{2l}+\ff{c_3}2\ll^{(3-4l)/(1-2l)}\Big]\\
&\quad\times \bigg( \E \exp \bigg[ \ll \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\bigg)^{1/2}.\end{split}\end{equation*}
As the argument works also for $t\land \tau_n$ in place of $t$, we may assume priorly that the left-hand side of the above inequality is finite, so that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*} \E \exp\bigg[\ll \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\le \exp\Big[2\ll (1+|x|^2+|y|^2)^{2l}+c_3\ll^{(3-4l)/(1-2l)}\Big].\end{equation*}
Letting $$\ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)= \ff{2c_1(t^2+t^{-4})}{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2} \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{(1+c_2)t},$$ and combining the above inequality with (\ref{EG}) and (\ref{si}), we arrive at
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation}\label{EXP2}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\Big(|\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|-\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\big\{P_t(f\log f)-(P_t f)\log P_t f\big\}\Big)(x,y)\\
&\qquad\le \ff {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_t f(x,y)}2\log \E\exp\bigg[\ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho) \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\\
&\qquad\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_t f(x,y)\Big\{ \ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho) (1+|x|^2+|y|^2)^{2l} +\ff {c_3} 2 \ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)^{(3-4l)/(1-2l)}\Big\}\\
&\qquad\le \ff {P_t f(x,y) \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{c(1+t)}} {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho t^4}\big\{(|x|^2+|y|^2)^{2l} + \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^{4(l-1)/(1-2l)}t^{8(l-1)/(1-2l)}\big\}\end{split}\end{equation}
for some constant $c>0$. This proves the desired estimate for $t\in (0,1]$, and hence for all $t>0$ as observed in the proof of Corollary \ref{C2}.
(3) Let {\bf (H)} hold for $l=\ff 1 2$, so that (\ref{EXP}) reduces to
$$\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s}\le
\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll (1+|x|^2+|y|^2)}\Big\{\E \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{4\ll^2\|\sigma} \def\ess{\text{\rm{ess}}\|^2 \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s}\Big\}^{1/2}.$$
Taking $\ll= (2\|\sigma} \def\ess{\text{\rm{ess}}\|)^{-2}$ we obtain
$$\E \exp\bigg[\ff 1 {4\|\sigma} \def\ess{\text{\rm{ess}}\|^2} \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\le
\exp\Big[\ff 1 {4\|\sigma} \def\ess{\text{\rm{ess}}\|^2} (1+|x|^2+|y|^2)\Big].$$ Obviously, there exists a constant $c>0$ such that if $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\ge t^{-2}\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{c(1+t)}$ then
$\ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)\le (2\|\sigma} \def\ess{\text{\rm{ess}}\|)^{-2}$ so that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\Big(|\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|-\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\big\{P_t(f\log f)-(P_t f)\log P_t f\big\}\Big)(x,y)\\
&\qquad\le \ff {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_t f(x,y)}2\log \bigg(\E\exp\bigg[\ff 1 {4\|\sigma} \def\ess{\text{\rm{ess}}\|^2} \int_0^t \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{-(1+c_2)s} (1+|X_s|^2+|Y_s|^2)^{2l}\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]\bigg)^{4\|\sigma} \def\ess{\text{\rm{ess}}\|^2/\ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)}\\
&\qquad\le \ff {\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho P_t f(x,y)}{2\ll_t(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho)}\big(1+|x|^2+|y|^2\big)\le\ff{c' P_tf(x,y)}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big( 1+|x|^2+|y|^2\big)\end{split}\end{equation*}holds for some constant $c'>0$.
\end{proof}
\paragraph{Example 3.1} (Kinetic Fokker-Planck equation)
Let us consider once again the Example 2.1 introduced previously, and remark that the result of Corollary \ref{C3} (1) holds without the first assumption in {\bf (H)}, so that we get a pointwise version of Villani \cite[Th. A.8]{Vil09} under the same type of condition (polynomial growth at most), and thus recover its $L_2$ bound (constants are however rather difficult to compare).
\subsection{A general case}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor}\label{C4} Assume {\bf (A)}. Then there exists a constant $c>0$ such that
\beq\label{A1} |\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|^2 \le c\Big(\ff 1 {(1\land t)^3}+ \ff W{1\land t}\Big) P_t f^2,\ \ f\in \B_b(\R^{m+d}).\end{equation}
If moreover there exist constants $\ll, K>0$ and a $C^2$-function $\tilde} \def\Ric{\text{\rm{Ric}} W\ge 1$ such that
\beq\label{A2} \ll W\le K-\ff{L\tilde} \def\Ric{\text{\rm{Ric}} W}{\tilde} \def\Ric{\text{\rm{Ric}} W},\end{equation} then there exist constants $c,\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_0>0$ such that
\beq\label{GGA} |\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P_t(f\log f)-(P_t f)\log P_t f\big\} + \ff{c}{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho}\Big\{ \ff 1 {(t\land 1)^3} +\ff{\log \tilde} \def\Ric{\text{\rm{Ric}} W}{(t\land 1)^2}\Big\} P_t f\end{equation} holds for
$f\in \B_b^+(\R^{m+d})$ and $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\ge \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_0/t$. \end{cor}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Again, it suffices to prove for $t\in (0,1].$ By (\ref{uv}) and taking $z\in A^{-1}h_1$ such that $|z|=\|A^{-1}\|\cdot |h_1|$, there exists a constant $c>0$ such that
$$|\LL(h,z,s)|\le \ff c {t^2} |h|,\ \ |\Theta(h,z,s)|\le \ff c t|h|.$$ So, by {\bf (A)}
\beq\label{AA0} \big|u''(s)z-v''(s)h_2+\nabla} \def\pp{\partial} \def\EE{\scr E_{\Theta(h,z,s)}Z(X_s,Y_s)\big|^2\le \ff c {t^4} + \ff c {t^2} W(X_s,Y_s)\end{equation} holds for some constant $c>0.$ Since $W\ge 1$ and $\E W(X_s,Y_s)\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{Cs}W$,
this and (\ref{3.1}) yield that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\nabla} \def\pp{\partial} \def\EE{\scr E P_t f|^2 &\le c_1 (P_t f^2)\bigg\{\int_0^t |\LL(h,z,s)|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s + \E\int_0^t |\Theta(h,z,s)|^2W(X_s,Y_s)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg\}\\
&\le c_2\Big(\ff 1 {t^3} +\ff W t\Big) P_t f^2 \end{split}\end{equation*} holds for some constants $c_1,c_2>0.$
Next, it is easy to see that the process
$$M_s:=\tilde} \def\Ric{\text{\rm{Ric}} W(X_s,Y_s) \exp\bigg[-\int_0^s \ff{L\tilde} \def\Ric{\text{\rm{Ric}} W}{\tilde} \def\Ric{\text{\rm{Ric}} W}(X_r,Y_r)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D r\bigg]$$ is a local martingale, and thus a supermartingale due to the Fatou lemma. Combining this with (\ref{A2}) and noting that $\tilde} \def\Ric{\text{\rm{Ric}} W\ge 1$, we obtain
\beq\label{*AA}\E\text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{\ll \int_0^t W(X_s, Y_s)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s} \le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{ K t}\E M_t\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{K t} \tilde} \def\Ric{\text{\rm{Ric}} W.\end{equation} Then the second assertion follows from (\ref{3.2}) and (\ref{AA0}) since for any constant $\aa>0$ there exists a constant $c_2>0$ such that for any $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho t\ge \ss{\aa/\ll}$,
$$\E\exp\bigg[\ff{\aa}{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2t^2} \int_0^tW(X_s,Y_s)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg]
\le \bigg(\E \exp\bigg[\ll \int_0^t W(X_s,Y_s)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s \bigg]\bigg)^{\aa/(\ll\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho^2t^2)}. $$
\end{proof}
\section{Harnack inequality and applications}
The aim of this section is to establish the log-Harnack inequality introduced in \cite{RW, W10} and the Harnack inequality with power due to \cite{W97}. Applications of these inequalities to heat kernel estimates as well as Entropy-cost inequalities can be found in e.g. \cite{RW,W10}. We first consider the general case with assumption {\bf (A)} then move to the more specific setting with assumption {\bf (H)}. Again, we only consider
the time-homogenous case.
\subsection{Harnack inequality under {\bf (A)}}
We first introduce a result, essentially due to \cite{ATW09}, that the entropy-gradient estimate (\ref{3.2}) implies the Harnack inequality with a power.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{prp}\label{P4.0} Let $\H$ be a Hilbert space and $P$ a Markov operator on $\B_b(\H).$ Let $h\in\H$ such that for some $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h\in (0,1)$ and measurable function $\gg_h: [\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h,\infty)\times \H\to (0,\infty)$,
\beq\label{EGF} |\nabla} \def\pp{\partial} \def\EE{\scr E_h Pf|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \big\{P(f\log f)- (Pf)\log Pf\big\} +\gg_h(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho,\cdot)Pf,\ \ \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\ge \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h\end{equation}holds for all positive $f\in\B_b(\H)$.
Then for any $\aa\ge \ff 1 {1-\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h}$ and positive $f\in\B_b(\H)$,
$$(Pf)^\aa(\x)\le Pf^\aa(\x+h) \exp\bigg[\int_0^1 \ff{\aa}{1+(\aa-1)s}\gg_h\Big(\ff{\aa-1}{1+(\aa-1)s}, \x+sh\Big)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s\bigg],\ \ \x\in\H.$$\end{prp}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Let $\bb(s)= 1+(\aa-1)s.$ We have $\ff{\aa-1}{\bb(s)}\ge \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h$ provided $\aa\ge \ff 1 {1-\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h}.$ Then
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\ff{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D}{\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s} \log (P f^{\bb(s)})^{\aa/\bb(s)}(\x+sh)\\
&=\ff{\aa(\aa-1)\{P (f^{\bb(s)}\log f^{\bb(s)})-(P f^{\bb(s)})\log P f^{\bb(s)}\}}{\bb(s)^2P f^{\bb(s)}}(\x+sh)
+\ff{\aa \nabla} \def\pp{\partial} \def\EE{\scr E_h P f^{\bb(s)}}{\bb(s)P f^{\bb(s)}}(\x+sh)\\
&\ge -\ff{\aa}{\bb(s)} \gg_h\Big(\ff{\aa-1}{\bb(s)},\ \x+sh\Big),\ \ \ s\in [0,1].\end{split}\end{equation*} Then the proof is completed by taking integral over $[0,1]$ w.r.t. $\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s$. \end{proof}
Below is a consequence of (\ref{GGA}) and Proposition \ref{P4.0}.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor}\label{C0} Let {\bf (A)} and $(\ref{A2})$ hold. Then there exist constants $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_0,c>0$ such that for any $\aa>1, t>0$ and positive $f\in \B_b(\R^{m+d})$,
\beq\label{HHH} (P_t f)^\aa(\x)\le P_t f^\aa(\x+h) \exp\bigg[ \ff{\aa c |h|^2}{\aa-1}\bigg(\ff 1 {(1\land t)^3}+\ff {\int_0^1 \log\tilde} \def\Ric{\text{\rm{Ric}} W(\x+sh)\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s} {(1\land t)^2}\bigg)\bigg]\end{equation} holds for $\x,h\in \R^{m+d}$ with $|h|<\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_0 t.$\end{cor}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} By (\ref{GGA}),
$$|\nabla} \def\pp{\partial} \def\EE{\scr E_h P_t f|\le \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho |h|\big\{P_t(f\log f)-(P_t f)\log P_t f\big\}
+ \ff{c}{\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho}\Big\{ \ff 1 {(t\land 1)^3} +\ff{\log \tilde} \def\Ric{\text{\rm{Ric}} W}{(t\land 1)^2}\Big\} P_t f$$ holds for $\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho\ge \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_0/t$. Thus, (\ref{EGF}) holds for $P=P_t$ and
$$\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho_h= \delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho _0 |h|/t,\ \ \ \gg_h(\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho,\x)= \ff{c|h|^2}\delta} \def\DD{\Delta} \def\vv{\varepsilon} \def\rr{\rho \Big(\ff 1 {t^3} +\ff{\log\tilde} \def\Ric{\text{\rm{Ric}} W(\x)}{t^2}\Big).$$ Therefore, the desired Harnack inequality follows from Proposition \ref{P4.0}.
\end{proof}
To derive the log-Harnack inequality, we need the following slightly stronger condition than the second one in {\bf (A)}: there exists an increasing function $U$ on $[0,\infty)$ such that
\beq\label{WFY} |Z(\x)-Z(\y)|^2\le |\x-\y|^2 \big\{U(|\x-\y|) + \ll W(\y)\big\},\ \ \x,\y\in \R^{m+d}.\end{equation}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thm}\label{T4.1} Assume {\bf (A)} such that $(\ref{WFY})$ holds. Then there exists a constant $c>0$ such that
$$P_t\log f(\x) - \log P_t f(\y)
\le c|\x-\y|^2\bigg\{\ff 1 {(1\land t)^3} + \ff {U((1\lor t^{-1})|\x-\y|) + W(\y)}{t\land 1}\bigg\}$$ holds for any $t>0,$ positive function $f\in \B_b(\R^{m+d})$, and $\x,\y \in\R^{m+d}.$
\end{thm}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} Again as in the proof of Corollary \ref{C2}, it suffices to prove for $t\in (0,1].$ Let $\x=(x,y)$ and $\y=(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)$. We will make use of the coupling constructed in the proof of Theorem \ref{T1.1} for $\vv=1, h= (x-\tilde} \def\Ric{\text{\rm{Ric}} x, y-\tilde} \def\Ric{\text{\rm{Ric}} y) $ and $(u,v)$ being in (\ref{uv1}). We have $(X_t,Y_t)=(X_t^1,Y_t^1)$, and $(X_s^1,Y_s^1)_{s\in [0,t]}$ is generated by $L$ under the probability $\Q_1= R_t^1\P$. So, by the Young inequality (see \cite[Lemma 2.4]{ATW09}), we have
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} P_t \log f(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y) &=\E \big(R_t^1\log f(X_t^1, Y_t^1)\big) = \E\big(R_t^1\log f(X_t,Y_t)\big)\\
&\le \E(R_t^1\log R_t^1)+ \log\E f(X_t,Y_t)= \log P_t f(x,y)+ \E(R_t^1\log R_t^1).\end{split}\end{equation*} Combining this with (\ref{NNm}) we arrive at
\beq\label{BC} P_t\log f(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y) - \log P_t f(x,y)\le \ff 1 2 \E_{\Q_1}\int_0^t |\sigma} \def\ess{\text{\rm{ess}}^{-1}\xi_s^1|^2\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D s.\end{equation}
Taking $z$ such that $|z|\le \|A^{-1}\|\cdot |h_1|$, we obtain from (\ref{2.2'}), (\ref{WFY}), (\ref{uv1}) and (\ref{uv2}) that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} |\sigma} \def\ess{\text{\rm{ess}}^{-1}\xi_s^1|^2 & \le \Big\{|\LL(h,z,s)|^2 +|\Theta(h,z,s)|^2\big(U(|\Theta(h,z,s)|)+\ll W(X_s^1,Y_s^1)\big)\Big\}\\
&\le c|h|^2\Big\{\ff 1 {t^4} + \ff { U(|h|/t) + W(X_s^1,Y_s^2)}{t^2}\Big\}.\end{split}\end{equation*}
Combining this with (\ref{BC}) and noting that $LW\le CW$ implies $\E_{\Q_1} W(X_s^1, Y_s^1)\le \text{\rm{e}}} \def\ua{\underline a} \def\OO{\Omega} \def\oo{\omega^{Cs}W(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)$ for $s\in [0,t]$, we complete the proof.
\end{proof}
We conclude this part, we come back to Example 2.1 for the kinetic Fokker-Planck equation.
\paragraph{Example 4.1} In Example 2.1 let e.g. $V(x)= (1+|x|^2)^l$. Then {\bf (A)} and (\ref{WFY}) holds for $W(x,y)= \exp[2V(x)+|y|^2]$ and
$U(r)= c r^{2[(2l-1)\lor 1]}$ for some constant $c>0.$ Therefore, Theorem \ref{T4.1} applies.
Next, for the gradient-entropy inequality (\ref{GGA}) and (\ref{HHH}), let us consider for simplicity that $m=d=1$ and $V(x)=x^3$:
\beq\label{kfp2} \begin} \def\beq{\begin{equation}} \def\F{\scr F{cases} \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D X_t= Y_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t,\\
\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D Y_t= \text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D B_t -(X_t)^3\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t-Y_t\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D t.\end{cases}\end{equation}
In this case we have $Z(x,y)=-x^3-y$, so that
$$|Z(x,y)-Z(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)|^2\le c(|x-\tilde} \def\Ric{\text{\rm{Ric}} x|^2+|y-\tilde} \def\Ric{\text{\rm{Ric}} y|^2)(1+x^4+\tilde} \def\Ric{\text{\rm{Ric}} x^4).$$ Next, let $W(x,y)=1+ \ff 1 2 x^4+ y^2$. We have
$$LW(x,y)= 2y x^3 + 1 -2 x^3 y -2 y^2 = 1- 2 y^2 \le W(x,y).$$ Thus, (\ref{A1}) holds for $U=0$. Moreover,
following the line of in \cite{Wu01,BCG08,DFG09}, consider $w(x,y)=a\big(\ff 1 2 x^4+y^2)+bxy$ for some well chosen constant $a,b$ and putting ${\tilde} \def\Ric{\text{\rm{Ric}} W}(x,y)=\exp(w-\inf w)$, we have
$$-\ff{L{\tilde} \def\Ric{\text{\rm{Ric}} W}}{\tilde} \def\Ric{\text{\rm{Ric}} W}\ge \aa W-K$$ for some constants $\aa,K>0$.
Indeed,
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split}
\ff{L \tilde W}{\tilde} \def\Ric{\text{\rm{Ric}} W}(x,y) &= L\log \tilde} \def\Ric{\text{\rm{Ric}} W(x,y)- \ff 1 2 |\pp_y \log \tilde} \def\Ric{\text{\rm{Ric}} W|^2(x,y)\\
&= a+2a^2y^2-2ax^3y-bx^4-2ay^2-bxy+2ax^3y+by\\
&\le a+(2a^2-2a+b(1+\vv/2))y^2-bx^4+bx^2/(2\vv)\\
&\le K-\aa (1+y^2+x^4)\end{split}\end{equation*}
holds for some constants $\aa,K>0$ by taking $a,b,\vv>0$ such that $2a^2-2a+b(1+\varepsilon/2)<0$.
Therefore, (\ref{A2}) holds for some $\ll, K>0$ so that (\ref{GGA}) and (\ref{HHH}) hold.
\subsection{Harnack inequality under assumption (H)}
As shown in \cite{ATW09}, the derivative estimate (\ref{1.4}) will enable us to prove an Harnack inequality with a power in the sense of \cite{W97}. More precisely, we have the following result.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thm} \label{THarnack} Let $(\ref{NZ})$ hold and let $\Phi_t$ be in $(\ref{Phi})$. Then for any $t>0,\aa>1$ and positive function $f\in \B_b(\R^{m+d}),$
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation} \label{Har}(P_t f)^\aa(x,y)\le (P_t f^\aa)(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\exp\Big[\ff{\aa}{\aa-1}\Phi_t(|x-\tilde} \def\Ric{\text{\rm{Ric}} x|,|y-\tilde} \def\Ric{\text{\rm{Ric}} y|)\Big],\ \ (x,y), (\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\in \R^{m+d}\end{equation}
holds. Consequently,
\beq\label{LHar} P_t \log f(x,y)\le \log P_t f(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y) + \Phi_t(|x-\tilde} \def\Ric{\text{\rm{Ric}} x|,|y-\tilde} \def\Ric{\text{\rm{Ric}} y|),\ \ (x,y), (\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\in \R^{m+d}.\end{equation}
\end{thm}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} It is easy to see that (\ref{Har}) follows from (\ref{1.4}) and Proposition \ref{P4.0}. Next, according to \cite[Proposition 2.2]{W10}, (\ref{LHar}) follows from $(\ref{Har})$ since $\R^{m+d}$ is a length space under the metric
$$\rr((x,y),(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)):= \ss{\Phi_t(|x-\tilde} \def\Ric{\text{\rm{Ric}} x|, |y-\tilde} \def\Ric{\text{\rm{Ric}} y|)}.$$ So, (\ref{Har}) implies (\ref{LHar}).
\end{proof}
The next result extends Theorem \ref{THarnack} to unbounded $\nabla} \def\pp{\partial} \def\EE{\scr E Z$.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thm} \label{THarnack2} Assume {\bf (H)}. Then there exists a constant $c>0$ such that for any $t >0$ and positive $f\in\B_b(\R^{m+d})$,
\beq\label{LHar'} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split} & P_t \log f(\y)- \log P_t f(\x)\\
& \le |\x-\y|^2 \Big\{\ff c {(1\land t)^3}
+ \ff{ c}{(1\land t)^{2l}}\big(1+|\x|+|\y|\big)^{4l}\Big\}\end{split}\end{equation} holds for $ \x,\y\in \R^{m+d}.$
If {\bf (H)} holds for some $l<\ff 1 2$, then there exists a constant $c>0$ such that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation} \label{Har'} \begin} \def\beq{\begin{equation}} \def\F{\scr F{split}&(P_t f)^\aa(\x)\le (P_t f^\aa)(\y) \\
&\times \exp\Big[\ff{\aa c|\x-\y|^2}{(\aa-1)(1\land t)^4}
\Big\{(|\x|\lor |\y|)^{4l} + \big((\aa-1)(1\land t)^2\big)^{4(l-1)/(1-2l)}\Big\}\Big]\end{split}\end{equation}
holds for all $t>0,\aa>1, \x,\y\in \R^{m+d}$ and positive $f\in \B_b(\R^{m+d}).$
\end{thm}
\begin} \def\beq{\begin{equation}} \def\F{\scr F{proof} (\ref{LHar'}) follows from Theorem \ref{T4.1} since in this case {\bf (A)} and (\ref{WFY}) hold for $W(\x)= (1+|\x|^2)^{2l} $ and $U(r)=c r^{2l}$ for some $\ll,c>0;$ while (\ref{Har'}) follows from Corollary \ref{C3}(2) and Proposition \ref{P4.0}.
\end{proof}
According to \cite[Proposition 2.4]{W10}, we have the following consequence of Theorems \ref{THarnack} and \ref{THarnack2}.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor} \label{C1.4} Let $p_t$ be the transition density of $P_t$ w.r.t. some $\sigma} \def\ess{\text{\rm{ess}}$-finite measure $\mu$ equivalent to the Lebesgue measure on $\R^{m+d}$. Let $\Phi_t$ be in Theorem $\ref{THarnack}$.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate} \item[$(1)$] $(\ref{NZ})$ implies
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{\R^{m+d}} \bigg(\ff{p_t((x,y),\z)}{p_t((\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y),\z)} \bigg)^{1/(\aa-1)} p_t((x,y),\z)
\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \z)\le\exp\bigg[\ff{\aa }{(\aa-1)^2}\Phi_t(|\tilde} \def\Ric{\text{\rm{Ric}} x-x|, |\tilde} \def\Ric{\text{\rm{Ric}} y-y|)\bigg],\\
&\int_{\R^{m+d}} p_t((x,y),\z)\log\ff{p_t((x,y),\z)}{p_t((\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y),\z)} \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\z)\le
\Phi_t(|\tilde} \def\Ric{\text{\rm{Ric}} x-x|, |\tilde} \def\Ric{\text{\rm{Ric}} y-y|).\end{split}\end{equation*} for any $t>0$ and $(x,y),(\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)\in \R^{m+d}.$
\item[$(2)$] If {\bf (H)} holds for some $l\in (0,\ff 1 2)$, then there exists a constant $c>0$ such that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{\R^{m+d}} \bigg(\ff{p_t(\x,\z)}{p_t(\y,\z)} \bigg)^{1/(\aa-1)} p_t(\x,\z)\mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \z)\\
&\le\exp\bigg[\ff{\aa c|\x-\y|^2}{(\aa-1)^2(1\land t)^4}
\Big\{(|\x|\lor |\y)^{4l} + \big((\aa-1)(1\land t)^2\big)^{4(l-1)/(1-2l)}\Big\}\bigg]\end{split}\end{equation*}
holds for all $t>0$ and $\x,\y\in \R^{m+d}.$
\item[$(3)$] If {\bf (H)} holds then there exists a constant $c>0$ such that
\begin} \def\beq{\begin{equation}} \def\F{\scr F{equation*}\begin} \def\beq{\begin{equation}} \def\F{\scr F{split} &\int_{\R^{m+d}} p_t(\x,\z)\log\ff{p_t(\x,\z)}{p_t(\y,\z)} \mu(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\z)\\
&\le
|\x-\y|^2 \Big\{\ff c {(1\land t)^3}
+ \ff{c}{(1\land t)^{2l}}\big(1+|\x|+|\y|\big)^{4l}\Big\}\end{split}\end{equation*} holds for all $t>0$ and $\x,\y\in \R^{m+d}.$ \end{enumerate} \end{cor}
Next, for two probability measures $\mu$ and $\nu$, let
$\scr C} \def\aaa{\mathbf{r}} \def\r{r(\nu,\mu)$ be the class of their couplings, i.e. $\pi\in \scr C} \def\aaa{\mathbf{r}} \def\r{r(\nu,\mu)$ if $\pi$ is a probability meadsure on $\R^{m+d}\times \R^{m+d}$ such that $\pi(\R^{m+d}\times \cdot)=\mu(\cdot)$ and $\pi(\cdot\times \R^{m+d})=\nu(\cdot)$.
Then according to the proof of \cite[Corollary 1.2(3)]{RW}, Theorems \ref{THarnack} and \ref{THarnack2} also imply the following entropy-cost inequalities. Recall that for any non-negative symmetric measurable function ${\bf c}$ on $\R^{m+d}\times \R^{m+d}$, and for any two probability measures $\mu,\nu$ on $\R^{m+d}$, we call
$$W_{\bf c}(\nu,\mu):= \inf_{\pi\in\scr C(\nu,\mu)}\int_{\R^{m+d}\times\R^{m+d}}{\bf c}(\x,\y)\,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D\pi(\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \x,\text{\rm{d}}} \def\bb{\beta} \def\aa{\alpha} \def\D{\scr D \y)$$ the transportation-cost between these two distributions induced by the cost function ${\bf c}$, where $\scr C(\nu,\mu)$ is the set of all couplings of $\nu$ and $\mu$.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{cor} \label{C4b} Let $P_t$ have an invariant probability measure $\mu$, and let $P^*$ be the adjoint operator of $P$ in $L^2(\mu).$ \begin} \def\beq{\begin{equation}} \def\F{\scr F{enumerate} \item[$(1)$] If $(\ref{NZ})$ holds then
\beq\label{TC} \mu(P_t^*f\log P_t^*f)\le W_{{\bf c}_t}(f\mu,\mu),\ \ t>0, f\ge 0,\mu(f)=1,\end{equation} where
${\bf c}_t(x,y;\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y)= \Phi_t(|\tilde} \def\Ric{\text{\rm{Ric}} x-x|, |\tilde} \def\Ric{\text{\rm{Ric}} y-y|).$
\item[$(2)$] If {\bf (H)} holds, then there exists $c>0$ such that
$(\ref{TC})$ holds for
$${\bf c}_t(\x, \y)= |\x-\y|^2 \Big\{\ff c {(1\land t)^3}
+ \ff{ c}{(1\land t)^{2l}}\big(1+|\x|+|\y|\big)^{4l}\Big\}.$$\end{enumerate}\end{cor}
\paragraph{Remark 4.1}
\begin{enumerate}
\item[(I)] Recall that the Pinsker inequality says that for any two probability measures $\mu,\nu$ on a measurable space, the total variation norm of $u-v$ is dominated by the square root of twice relative entropy of $\nu$ w.r.t. $\mu$. Combining this inequality with (1) of Corollary \ref{C1.4}, assuming thus $\|\nabla Z\|_\infty<\infty$, we get
$$\|P_t((x,y),\cdot)-P_t((\tilde} \def\Ric{\text{\rm{Ric}} x,\tilde} \def\Ric{\text{\rm{Ric}} y),\cdot)\|_{TV}\le \sqrt{2\,\Phi_t(|\tilde} \def\Ric{\text{\rm{Ric}} x-x|, |\tilde} \def\Ric{\text{\rm{Ric}} y-y|)},$$
which may be useful as an alternative to small set evaluation in Meyn-Tweedie's approach for convergence to equilibrium for the kinetic Fokker-Planck equation.
\item[(II)] Using Villani's result \cite[Th.39]{Vil09} in the kinetic Fokker Planck case which asserts that if $|\nabla^2V|$ is bounded and $\mu$ as a product measure satisfies a logarithmic Sobolev inequality, then there is an exponential convergence towards equilibrium in entropy, so that
$$\mu(P_s^*f\log P_s^*f)\le Ce^{-Ks}\mu(P_1^*f\log P_1^*f),\ \ f\ge 0, \mu(f)=1, s\ge 1$$
holds for some constant $C>0$. Combining this with Talagrand inequality implied by the logarithmic Sobolev inequality (see \cite{OV00}) and using Corollary \ref{C4b}(1), we get
$$W_2^2(P_sf\mu,\mu)\le C'e^{-Ks}W_2^2(f\mu,\mu),\ \ f\ge 0, \mu(f)=1, s\ge 1$$ for some constant $C'>0$,
where $W_2^2= W_{\bf c}$ for ${\bf c}(\x;\y):= |\x-\y|^2.$ This
generalizes the exponential convergence in Wasserstein distance derived in \cite{BGM10} for the non interacting case.
\end{enumerate}
\paragraph{Acknowledgement} The second named author would like to thank Professor Xicheng Zhang for introducing his very interesting paper
\cite{Zhang}. Both authors would like to thank the referee for helpful comments.
\begin} \def\beq{\begin{equation}} \def\F{\scr F{thebibliography}{99}
\bibitem{AT99} M. Arnaudon, A. Thalmaier, \emph{Bismut type differentiation of semigroups,} Probability Theory and Mathematical Statistics 23--32, VSP/TEV, Utvecht and Viluius, 1999.
\bibitem{ATW06} M. Arnaudon, A. Thalmaier, F.-Y. Wang,
\emph{Harnack inequality and heat kernel estimates
on manifolds with curvature unbounded below,} Bull. Sci. Math. 130(2006), 223--233.
\bibitem{ATW09} M. Arnaudon, A. Thalmaier, F.-Y. Wang,
\emph{Gradient estimates and Harnack inequalities on non-compact Riemannian manifolds,}
Stoch. Proc. Appl. 119(2009), 3653--3670.
\bibitem{BCG08} D. Bakry, P. Cattiaux, A. Guillin, \emph{Rate of convergence for ergodic continuous Markov processes : Lyapunov versus Poincare,}
J. Func. Anal. 254 (2008), 727--759.
\bibitem{Bismut} J. M. Bismut, \emph{Large Deviations and the
Malliavin Calculus,} Boston: Birkh\"auser, MA, 1984.
\bibitem{BGM10} F. Bolley, A. Guillin, F. Malrieu, \emph{Trend to equlibrium and particle approximation for a weakly selfconsistent Vlasov-Fokker-Planck equation}, M2AN 44(5) (2010), 867--884.
\bibitem{DRW09} G. Da Prato, M. R\"ockner, F.-Y. Wang, \emph{Singular stochastic equations on Hilbert
spaces: Harnack inequalities for their transition semigroups,} J. Funct. Anal. 257 (2009), 992--017.
\bibitem{DFG09} R. Douc, G. Fort, A. Guillin, \emph{Subgeometric rates of convergence of f-ergodic strong Markov processes},
Stoch. Proc. Appl. 119 (2009) 897--923.
\bibitem{EL} K.D. Elworthy, Xue-Mei Li, \emph{Formulae for the
derivatives of heat semigroups,} J. Funct. Anal. 125(1994),
252--286.
\bibitem{ES} A. Es-Sarhir, M.-K. v. Renesse, M. Scheutzow,
\emph{Harnack inequality for functional SDEs with bounded memory,}
Electron. Commun. Probab. 14 (2009), 560--565.
\bibitem{K} H. Kawabi, \emph{The parabolic Harnack inequality for
the time dependent Ginzburg-Landau type SPDE and its application,}
Pot. Anal. 22(2005), 61--84.
\bibitem{LW} W. Liu, F.-Y. Wang, \emph{Harnack inequality and strong Feller
property for stochastic fast diffusion equations,} J. Math. Anal. Appl.
342(2008), 651--662.
\bibitem{OV00} F. Otto, C. Villani, \emph{Generalization of an inequality by Talagrand and links with the logarithmic
Sobolev inequality}, J. Funct. Anal., 173 (2000), 361--400.
\bibitem{Ouyang} S.-X. Ouyang, \emph{Harnack inequalities and applications for multivalued stochastic evolution equations,} Infin. Dimens. Anal. Quant. Probab. Relat. Topics 14(2011), 261--278.
\bibitem{ORW} S.-X. Ouyang, M. R\"ockner, F.-Y. Wang,
\emph{Harnack inequalities and applications for Ornstein-Uhlenbeck semigroups with
jump,} Pot. Anal. 36(2012), 301--315.
\bibitem{RW} M. R\"ockner, F.-Y. Wang, \emph{Log-Harnack inequality for stochastic differential equations in Hilbert spaces and its consequences, } Infin. Dimens. Anal. Quant. Probab. Relat. Topics 13(2010), 27--37.
\bibitem{Vil09} C. Villani, \emph{Hypocoercivity}. Mem. Amer. Math. Soc. 202 (2009), no. 950.
\bibitem{W97} F.-Y. Wang, \emph{Logarithmic Sobolev
inequalities on noncompact Riemannian manifolds,} Probability
Theory Relat. Fields 109(1997), 417--424.
\bibitem{W07} F.-Y. Wang, \emph{Harnack inequality and applications
for stochastic generalized porous media equations,} Ann. Probab. 35(2007),
1333--1350.
\bibitem{W10} F.-Y. Wang, \emph{Harnack inequalities on manifolds with boundary and applications,} J. Math. Pures Appl. 94(2010), 304--321.
\bibitem{W10b} F.-Y. Wang, \emph{Coupling and its applications}, Preprint, accessible on arXiv:1012.5687.
\bibitem{WX10} F.-Y. Wang, L. Xu, \emph{Derivative formula and applications for hyperdissipative stochastic Navier-Stokes/Burgers equations,} to appear in Infin. Dimens. Anal. Quant. Probab. Relat. Topics, accessible on arXiv:1009.1464.
\bibitem{WWX10} F.-Y. Wang, J.-L. Wu, L. Xu, \emph{Log-Harnack inequality for stochastic Burgers equations and applications,} J. Math. Anal. Appl. 384(2011), 151--159.
\bibitem{WY10} F.-Y. Wang, C. Yuan, \emph{Harnack inequalities for functional SDEs with multiplicative noise and applications}, Stoch. Proc. Appl. 121(2011), 2692--1710.
\bibitem{Wu01} L. Wu, \emph{Large and moderate deviations and exponential convergence for stochastic damping Hamiltonian systems}, Stoch. Proc. Appl., 91 (2001), 205--238.
\bibitem{Z} T.-S. Zhang, \emph{White noise driven SPDEs with reflection: strong Feller properties and Harnack inequalities,} Potential Anal.
33 (2010),137--151.
\bibitem{Zhang} X. Cheng, \emph{Stochastic flows and Bismut formulas for stochastic Hamiltonian systems,} Stoch. Proc. Appl. 120(2010), 1929--1949.
\end{thebibliography}
\end{document} |
\section{Introduction}
We refer to~\cite{west} for graphical notations and terminologies
not described in this paper. Throughout the paper, $G=(V,E)$ is a
finite, simple, and connected graph. The distance between two
vertices $u$ and $v$, denoted by $d(u,v)$, is the length of a
shortest path between $u$ and $v$ in $G$. Also, $N(v)$ is the set
of all neighbors of vertex $v$ and $deg(v)=|N(v)|$ is the degree
of vertex $v$. The maximum degree of the graph $G$, $\Delta(G)$ is
$max_{v\in V(G)}deg(v)$. We mean by $\omega(G)$, the number of
vertices in a maximum clique in $G$. For a subset $S$ of $V(G)$,
$G\setminus S$ is the induced subgraph $\langle V(G)\setminus
S\rangle$ by $V(G)\setminus S$ of $G$. A set $S\subseteq V(G)$ is
a separating set in $G$ if $G\setminus S$ has at least two
connected components. We call a vertex $v\in V(G)$ a cut vertex of
$G$ if $\{v\}$ is a separating set in $G$. If $G\neq K_n$ has no
cut vertex, then $G$ is called a $2$-connected graph. The
notations $u\sim v$ and $u\nsim v$ denote the adjacency and
non-adjacency relation between $u$ and $v$, respectively. The
symbol $(v_1,v_2,\ldots, v_n)$ represents a path of order $n$,
$P_n$.
For an ordered set $W=\{w_1,w_2,\ldots,w_k\}\subseteq V(G)$ and a
vertex $v$ of $G$, the $k$-vector
$$r(v|W):=(d(v,w_1),d(v,w_2),\ldots,d(v,w_k))$$
is called the ({\it metric}) {\it representation} of $v$ with
respect to $W$. The set $W$ is called a {\it resolving set} for
$G$ if distinct vertices have different representations. A
resolving set for $G$ with minimum cardinality is called a {\it
basis} of $G$, and its cardinality is the {\it metric dimension}
of $G$, denoted by $\beta(G)$.
For example, the graphs $G$ and $H$
in Figure~\ref{fig} have the basis $B=\{v_1,v_2\}$ and hence
$\beta(G)=\beta(H)=2$. The representations of vertices of $G$ with
respect to $B$ are
$$r(v_1|B)=(0,1),\quad r(v_2|B)=(1,0),\quad r(v_3|B)=(2,1),\quad
r(v_4|B)=(2,2),\quad r(v_5|B)=(1,2).$$ Also, the representations
of vertices of $H$ with respect to $B$ are
$$r(v_1|B)=(0,1),\quad r(v_2|B)=(1,0),\quad r(v_3|B)=(1,1),\quad
r(v_4|B)=(2,2),\quad r(v_5|B)=(1,2).$$
\begin{figure}[ht]
\centering \unitlength=.4mm
\caption{\label{fig} \scriptsize{$bas(G)=\beta(G)=res(G)$ and
$bas(H)\neq\beta(H)\neq res(H)$.}}
\end{figure}
\\To see that whether a given set $W$ is a resolving set for $G$,
it is sufficient to look at the representations of vertices in
$V(G)\backslash W$, because $w\in W$ is the unique vertex of $G$
for which $d(w,w)=0$. When $W$ is a resolving set for $G$, we say
that $W$ {\it resolves} $G$. In general, we say an ordered set $W$
resolves a set $T$ of vertices in $G$, if the representations of
vertices in $T$ are distinct with respect to $W$. When $W=\{x\}$,
we say that vertex $x$ resolves $T$.
\par In~\cite{Slater1975}, Slater introduced the idea of a resolving
set and used a {\it locating set} and the {\it location number}
for what we call a resolving set and the metric dimension,
respectively. He described the usefulness of these concepts when
working with U.S. Sonar and Coast Guard Loran stations.
Independently, Harary and Melter~\cite{Harary} discovered the
concept of the location number as well and called it the metric
dimension. For more results related to these concepts
see~\cite{cartesian product,bounds,sur1,landmarks,sur2}. The
concept of a resolving set has various applications in diverse
areas including coin weighing problems~\cite{coin}, network
discovery and verification~\cite{net2}, robot
navigation~\cite{landmarks}, mastermind game~\cite{cartesian
product}, problems of pattern recognition and image
processing~\cite{digital}, and combinatorial search and
optimization~\cite{coin}.
\par The following simple result is very useful.
\begin{obs}~\rm\cite{extermal}\label{twins}
Let $G$ be a graph and $u,v\in V(G)$ such that,
$N(v)\backslash\{u\}=N(u)\backslash\{v\}$. If $W$ resolves $G$,
then $u$ or $v$ is in $W$.
\end{obs}
It is obvious that for a graph $G$ of order $n$,
$1\leq\beta(G)\leq n-1$.
\begin{lem}~\rm\cite{Ollerman}\label{B=1,B=n-1}
Let $G$ be a graph of order $n$. Then,\begin{description}\item (i)
$\beta(G)=1$ if and only if $G=P_n$,\item (ii) $\beta(G)=n-1$ if
and only if $G=K_n$.
\end{description}
\end{lem}
\par The {\it basis number}, $bas(G)$, of $G$ is the maximum integer $r$
such that, every $r$-set of vertices of $G$ is a subset of some
basis of $G$. Also, the {\it resolving number}, $res(G)$, of $G$
is the minimum integer $k$ such that, every $k$-set of vertices of
$G$ is a resolving set for $G$. These parameters are introduced
in~\cite{basis} and ~\cite{res(G)}, respectively. Clearly, if $G$
is a graph of order $n$, then $0\leq bas(G)\leq \beta(G)$ and
$\beta(G)\leq res(G)\leq n-1$. Chartrand et al. in~\cite{basis}
considered graphs $G$ with $bas(G)=\beta(G)$. They called these
graphs {\it randomly $k$-dimensional} graphs, where $k=\beta(G)$.
Obviously, $bas(G)=\beta(G)$ if and only if $res(G)=\beta(G)$. In
the other word, a randomly $k$-dimensional graph is a graph which
every $k$-set of its vertices is a basis. For example in graph $G$
of Figure~\ref{fig}, if $W$ is a set of two adjacent vertices,
then the representations of vertices in $V(G)\setminus W$ with
respect to $W$ are $(1,2),~(2,2)$, and $(2,1)$. Also, if $W$ is a
set of two non-adjacent vertices, then the representations of
vertices in $V(G)\setminus W$ with respect to $W$ are
$(1,1),~(1,2)$, and $(2,1)$. Therefore, $G$ is a randomly
$2$-dimensional graph. But, in graph $H$ of Figure~\ref{fig},
$\{v_1,v_4\}$ is not a resolving set, hence $H$ is not a randomly
$2$-dimensional graph. Since $\{v_1,v_2\}$, $\{v_1,v_3\}$, and
$\{v_4,v_5\}$ are bases of $H$, $bas(H)=1$. Also, $res(H)=3$,
because every $3$-set of $V(H)$ is a resolving set in $H$.
\par Obviously, $K_1$ and $K_2$ are the only
randomly $1$-dimensional graphs. Chartrand et al.~\cite{basis}
proved that a graph $G$ is randomly $2$-dimensional if and only if
$G$ is an odd cycle. In this paper, we first characterize all
graphs of order $n$ and resolving number $1$ and
$n-1$. Then, we provide
some properties of randomly $k$-dimensional graphs.
\section{Main Results}\label{main}
We first characterize all graphs $G$ with $res(G)=1$ and all
graphs $G$ of order $n$ with $res(G)=n-1$.
\begin{thm} \label{res=1,res=n-1}
Let $G$ be a graph of order $n$. Then,
\begin{description} \item (i) $res(G)=1$ if and only if
$G\in\{P_1,P_2\}$.\item (ii) $res(G)=n-1$ if and only if
$N(v)\backslash\{u\}=N(u)\backslash\{v\}$, for some $u,v\in V(G)$.
\end{description}
\end{thm}
\begin{proof}{ (i) It is easy to see that for $G\in
\{P_1,P_2\}$, $res(G)=1$. Conversely, let $res(G)=1$. Thus, $1\leq
\beta(G)\leq res(G)=1$ and hence, $\beta(G)=1$. Therefore, by
Theorem~\ref{B=1,B=n-1}, $G=P_n$. If $n\geq3$, then $P_n$ has a
vertex of degree $2$ and this vertex does not resolve its
neighbors. Thus, $res(G)\geq 2$, which is a contradiction.
Consequently, $n\leq2$, that is $G\in\{P_1,P_2\}$.\\
(ii) Let $u,v\in V(G)$ such that,
$N(v)\backslash\{u\}=N(u)\backslash\{v\}$. If $res(G)\leq n-2$,
then the set $V(G)\setminus \{u,v\}$ is a resolving set for $G$.
But, by Observation~\ref{twins}, every resolving set for $G$
contains at least one of the vertices $u$ and $v$. This
contradiction implies that, $res(G)=n-1$. Conversely, let
$res(G)=n-1$. Thus, there exists a subset $T$ of $V(G)$ with
cardinality $n-2$ such that, $T$ is not a resolving set for $G$.
Assume that, $T=V(G)\setminus \{u,v\}$. If
$N(u)\backslash\{v\}\neq N(v)\backslash\{u\}$, then there exists a
vertex $w\in T$ which is adjacent to only one of the vertices $u$
and $v$ and hence, $d(u,w)\neq d(v,w)$. Since $w\in T$, $T$
resolves $G$, which is a contradiction. Therefore,
$N(u)\backslash\{v\}=N(v)\backslash\{u\}$.}\end{proof}
\begin{cor}\label{not twin} If $G\neq K_n$ is a randomly
$k$-dimensional graph, then for each pair of vertices $u,v\in
V(G)$, $N(v)\backslash\{u\}\neq N(u)\backslash\{v\}$.
\end{cor}
\begin{proof}{
If $N(v)\backslash\{u\}=N(u)\backslash\{v\}$ for some $u,v\in
V(G)$, then by Theorem~\ref{res=1,res=n-1}, $res(G)=n-1$, where
$n$ is the order of $G$. Since $G$ is a randomly $k$-dimensional
graph, $\beta(G)=res(G)=n-1$. Therefore, by
Theorem~\ref{B=1,B=n-1}, $G=K_n$, which is a contradiction. Hence,
for each $u,v\in V(G)$, $N(v)\backslash\{u\}\neq
N(u)\backslash\{v\}$. }\end{proof}
\begin{lemma}\label{delta>2} If $G$ is a randomly
$k$-dimensional graph with $k\geq2$ and minimum degree $\delta$,
then $\delta\geq2$.
\end{lemma}
\begin{proof}{ Suppose on the contrary that, there exists a vertex
$u\in V(G)$ with $deg(u)=1$. Let $v$ be the unique neighbor of $u$
and $T\subseteq V(G)$ be a subset of $V(G)$ with $|T|=k$ and
$u,v\in T$. Since $G$ is a randomly $k$-dimensional graph,
$T\setminus \{v\}$ is not a resolving set for $G$. Thus, there
exists a pair of vertices $x,y\in V(G)$ such that, $d(x,v)\neq
d(y,v)$ and $d(x,t)=d(y,t)$, for each $t\in T\setminus \{v\}$.
Hence, $d(x,u)=d(y,u)$. Clearly, if $u\in\{x,y\}$, then
$d(x,u)\neq d(y,u)$, which is a contradiction. Consequently,
$u\notin\{x,y\}$. Therefore, $d(x,u)=d(x,v)+1$ and
$d(y,u)=d(y,v)+1$. Thus, $d(x,v)=d(y,v)$. This contradiction
implies that $\delta\geq2$. }\end{proof}
\begin{thm}\label{2connected} If $k\geq2$, then every randomly $k$-dimensional graph
is $2$-connected.
\end{thm}
\begin{proof}{Suppose on the contrary that $u$ is a cut vertex in
$G$. Let $G_1$ be a connected component of $G\setminus \{u\}$. Set
$H_2:=G\setminus V(G_1)$ and $H_1:=\langle
V(G_1)\cup\{u\}\rangle$, the induced subgraph by $V(G_1)\cup\{u\}$
of $G$. Note that, for each $x\in V(H_1)$ and each $y\in V(H_2)$,
$d(x,y)=d(x,u)+d(u,y)$. By Lemma~\ref{delta>2}, $G$ does not have
any vertex of degree $1$. Therefore, $|V(H_1)|\geq3$ and
$|V(H_2)|\geq3$. Let $a,b\in V(H_2)$ and $V(H_1)$ resolves
$\{a,b\}$. Then, there exists a vertex $w\in V(H_1)$ such that,
$d(a,w)\neq d(b,w)$. Thus, $d(a,u)+d(u,w)\neq d(b,u)+d(u,w)$, that
is $d(a,u)\neq d(b,u)$. Hence, $V(H_1)$ resolves a pair of
vertices of $V(H_2)$ if and only if $u$ resolves this pair. If
$V(H_1)$ is a resolving set for $G$, then $\{u\}$ is a resolving
set for $H_2$. Therefore, by Theorem~\ref{B=1,B=n-1}, $H_2$ is a
path. Since $|V(H_2)|\geq3$, $G$ has a vertex of degree $1$, which
contradicts Lemma~\ref{delta>2}. Hence, $\beta(H_2)\geq2$ and
$V(H_1)$ does not resolve $G$. Now,
one of the following two cases can be happened.\vspace{3mm}\\
1. $u$ belongs to a basis of $H_2$. In this case $u$ along with
$\beta(H_2)-1$ vertices of $V(H_2)\setminus \{u\}$ forms a basis
$T$ of $H_2$. Since $\beta(H_2)\geq2$, there exists a vertex $x\in
T\setminus\{u\}$. Note that, $T\cup V(H_1)\setminus\{x\}$ is not a
resolving set for $G$, otherwise $T\setminus\{x\}$ is a resolving
set for $H_2$ of size $\beta(H_2)-1$. Thus,
$$res(G)\geq |T\cup V(H_1)|=\beta(H_2)+|V(H_1)|-1.$$ Now, let
$z\in V(G_1)$. Since $|V(H_1)|\geq3$ and $G_1$ is a connected
component of $G\setminus\{u\}$, $z$ has a neighbor in $G_1$, say
$v$. Therefore, $d(z,v)=1\neq d(y,v)$ for each $y\in
V(H_2)\setminus\{u\}$. Hence, the set $T\cup V(H_1)\setminus\{z\}$
is a resolving set for $G$. Thus, $$\beta(G)\leq |T\cup
V(H_1)\setminus\{z\}|= \beta(H_2)+|V(H_1)|-2.$$ Consequently,
$\beta(G)<res(G)$, which is a contradiction.
\vspace{3mm}\\
2. $u$ does not belong to any basis of $H_2$. Let $T$ be a basis
of $G$ and $x\in T$. Therefore, $T\cup V(H_1)\setminus\{x\}$ is
not a resolving set for $G$. Hence,
$$res(G)\geq |T\cup V(H_1)|=\beta(H_2)+|V(H_1)|.$$ Now, let
$z\in V(G_1)$. Similar to the previous case, $T\cup
V(H_1)\setminus\{z\}$ is a resolving set for $G$. Thus,
$$\beta(G)\leq |T\cup V(H_1)\setminus\{z\}|=
\beta(H_2)+|V(H_1)|-1.$$ Therefore, $\beta(G)<res(G)$, which is
impossible.
\par Consequently, $G$ does not have any cut vertex.
}\end{proof}
\begin{thm}\label{deg2 not adjacent} If $G$ is a randomly $k$-dimensional
graph with $k\geq4$, then there are no adjacent vertices of degree
$2$ in $G$.
\end{thm}
\begin{proof}{ Suppose on the contrary that $G$ has adjacent
vertices of degree $2$. Therefore, there is an induced subgraph
$P_r=(a_1,a_2,\ldots,a_r),~r\geq2$, such that, for each $i$,
$1\leq i\leq r$, $deg(a_i)=2$ in $G$. Let $x,y\in V(G)\setminus
V(P_r)$ and $x\sim a_1$, $y\sim a_r$. Since $k\geq4$, $G$ is not a
cycle. Thus, Theorem~\ref{2connected} implies that $x\neq y$,
otherwise, $x=y$ is a cut vertex in $G$. By assumption, $G$ has a
basis $B=\{x,y,a_i,a_j\}\cup T$, where $1\leq i\neq j\leq r$ and
$T$ is a subset of $V(G)\setminus \{x,y,a_i,a_j\}$ with $|T|=k-4$.
Now, one of the following cases can be happened.
\vspace{3mm}\\
1. $r$ is odd. Let
$B_1=B\cup\{a_{r+1\over2}\}\setminus\{a_i,a_j\}$. We claim that,
$B_1$ is a resolving set for $G$. Otherwise, there exist vertices
$u,v\in V(G)$ with $r(u|B_1)=r(v|B_1)$. If $v\in V(P_r)$ and
$u\notin V(P_r)$, then $d(v,a_{r+1\over2})\leq {r-1\over2}$ and
$d(u,a_{r+1\over2})\geq {r+1\over2}$. Hence, $r(u|B_1)\neq
r(v|B_1)$, which is a contradiction. Therefore, both of vertices
$u$ and $v$ belong to $V(P_r)$ or $V(G)\setminus V(P_r)$. If
$u,v\in V(P_r)$, then, $d(u,a_{r+1\over2})=d(v,a_{r+1\over2})$
implies $u,v\in\{a_{{r+1\over2}-i},a_{{r+1\over2}+i}\}$ for some
$i$, $1\leq i\leq{r-1\over2}$. On the other hand,
$d(x,a_{{r+1\over2}-i})={r+1\over2}-i$ and
$d(x,a_{{r+1\over2}+i})=\min\{{r+1\over2}+i,{r+1\over2}-i+d(x,y)\}$.
If ${r+1\over2}+i\leq{r+1\over2}-i+d(x,y)$, then
$d(x,a_{{r+1\over2}-i})\neq d(x,a_{{r+1\over2}+i})$, which is a
contradiction. Thus, ${r+1\over2}-i+d(x,y)<{r+1\over2}+i$ and
hence, ${r+1\over2}-i+d(x,y)={r+1\over2}-i$, because
$d(x,a_{{r+1\over2}-i})=d(x,a_{{r+1\over2}+i})$. Therefore,
$d(x,y)=0$, which contradicts $x\neq y$. Thus, $u,v\in
V(G)\setminus V(P_r)$. Since $r(u|B_1)=r(v|B_1)$ and $B$ is a
resolving set for $G$, there exists a vertex in $B\setminus
B_1=\{a_i,a_j\}\setminus\{a_{r+1\over2}\}$ which resolves
$\{u,v\}$. By symmetry, we can assume $a_i$ resolves $\{u,v\}$.
Therefore, $d(u,a_i)\neq d(v,a_i)$, $d(u,x)=d(v,x)$, and
$d(u,y)=d(v,y)$. But,
$$d(u,a_i)=\min\{d(u,x)+d(x,a_i),d(u,y)+d(y,a_i)\},$$and
$$d(v,a_i)=\min\{d(v,x)+d(x,a_i),d(v,y)+d(y,a_i)\}.$$If
$d(u,x)+d(x,a_i)\leq d(u,y)+d(y,a_i)$ and $d(v,x)+d(x,a_i)\leq
d(v,y)+d(y,a_i)$, then $d(u,x)+d(x,a_i)\neq d(v,x)+d(x,a_i)$,
which implies $d(u,x)\neq d(v,x)$, a contradiction. Similarly, if
$d(u,y)+d(y,a_i)\leq d(u,x)+d(x,a_i)$ and $d(v,y)+d(y,a_i)\leq
d(v,x)+d(x,a_i)$, then $d(u,y)\neq d(v,y)$, which is a
contradiction. Therefore, by symmetry, we can assume
$d(u,x)+d(x,a_i)\leq d(u,y)+d(y,a_i)$ and $d(v,y)+d(y,a_i)\leq
d(v,x)+d(x,a_i)$.
Thus,$$d(u,a_i)=d(u,x)+d(x,a_i)=d(v,x)+d(x,a_i)\geq d(v,a_i),$$
and$$d(v,a_i)=d(v,y)+d(y,a_i)=d(u,y)+d(y,a_i)\geq d(u,a_i).$$
These imply that $d(u,a_i)=d(v,a_i)$, which is a contradiction.
Therefore, $B_1$ is a resolving set for $G$ with cardinality
$k-1$.\vspace{3mm}\\
2. $r$ is even. Let
$B_2=B\cup\{a_{r\over2}\}\setminus\{a_i,a_j\}$. Similar to the
previous case, $B_2$ is a resolving set for $G$ with cardinality
$k-1$.
\par In both cases, we get a contradiction to the assumption
that $G$ is a randomly $k$-dimensional graph. Therefore, there are
no adjacent vertices of degree $2$ in $G$. }\end{proof}
\begin{thm}\label{|T|=k-1} If $G$ is a randomly $k$-dimensional
graph and $T$ is a separating set of $G$ with $|T|=k-1$, then
$G\setminus T$ has exactly two connected components and for each
pair of vertices $u,v\in V(G)\setminus T$ with $r(u|T)=r(v|T)$,
$u$ and $v$ belong to different components.
\end{thm}
\begin{proof}{ Since $\beta(G)=k$ and $|T|=k-1$, there exist two
vertices $u,v\in V(G)\setminus T$ with $r(u|T)=r(v|T)$. Let $H$ be
a connected component of $G\setminus T$ for which $u\notin H$ and
$v\notin H$. If $w\in H$, then there exist two vertices $s,t\in T$
such that, $d(u,w)=d(u,s)+d(s,w)$ and $d(v,w)=d(v,t)+d(t,w)$.
Since $r(u|T)=r(v|T)$, we have $d(u,s)=d(v,s)$ and
$d(u,t)=d(v,t)$. Therefore,
$$d(u,w)=d(u,s)+d(s,w)=d(v,s)+d(s,w)\geq d(v,w).$$ And
$$d(v,w)=d(v,t)+d(t,w)=d(u,t)+d(t,w)\geq d(u,w).$$ Hence,
$d(u,w)=d(v,w)$. Thus, $r(u|T\cup\{w\})=r(v|T\cup\{w\})$.
Consequently, $T\cup\{w\}$ is not a resolving set for $G$ and
$|T\cup\{w\}|=k$. This contradicts the assumption that $G$ is
randomly $k$-dimensional. Therefore, $G\setminus T$ has exactly
two components and $u$ and $v$ belong to different components.
}\end{proof}
\begin{cor}\label{Delta>k}
If $G$ is a randomly $k$-dimensional graph with $k\geq2$, then
$\Delta(G)\geq k$.
\end{cor}
\begin{proof}{ If $G=K_n$, then $\Delta(G)=n-1=k$. Now let $G\neq
K_n$. Suppose on the contrary that $\Delta(G)\leq k-1$. Let $u\in
V(G)$, $deg(u)=\Delta(G)$, and $T$ be a subset of $V(G)$ with
$|T|=k-1$ and $N(u)\subseteq T$. By Theorem~\ref{|T|=k-1},
$G\setminus T$ has exactly two connected components, of which one
of them is $\{u\}$. Since $|T|=k-1$ and $\beta(G)=k$, there exist
two vertices $x,y\in V(G)\setminus T$ such that, $r(x|T)=r(y|T)$.
By Theorem~\ref{|T|=k-1}, $x$ and $y$ belong to different
components. Therefore, one of them is $u$, say $x=u$. Since
$r(u|T)=r(y|T)$, we have $N(u)\subseteq N(y)$. By
Corollary~\ref{not twin}, $G$ does not have any pair of vertices
$u,v$ with $N(u)\setminus\{v\}=N(v)\setminus\{u\}$. Hence,
$N(u)\subset N(y)$, this contradicts $deg(u)=\Delta(G)$.
Therefore, $\Delta(G)\geq k$. }\end{proof}
\begin{cor}\label{deg not adjacent>k}
If $u$ and $v$ are two non-adjacent vertices in a randomly
$k$-dimensional graph, then $deg(u)+deg(v)\geq k$.
\end{cor}
\begin{proof}{ If $|N(u)\cup N(v)|\leq k-1$, then let $T$ be a subset of
$V(G)\setminus\{u,v\}$ with $|T|=k-1$ and $N(u)\cup N(v)\subseteq
T$. By Theorem~\ref{|T|=k-1}, $G\setminus T$ has exactly two
connected components $\{u\}$ and $\{v\}$. Hence, $|T|=n-2$. This
implies that $k=n-1$ and by Theorem~\ref{res=1,res=n-1}, $G=K_n$.
Consequently, $u\sim v$, which is a contradiction. Thus,
$deg(u)+deg(v)\geq|N(u)\cup N(v)|\geq k$.
}\end{proof}
\begin{thm}\label{omega<k+1} If $G$ is a randomly $k$-dimensional
graph of order at least $2$, then $\omega(G)\leq k+1$. Moreover, $\omega(G)=k+1$ if and
only if $G=K_n$.
\end{thm}
\begin{proof}{ Let $H$ be a clique of size $\omega(G)$ in $G$ and
$T$ be a subset of $V(H)$ with $|T|=\omega(G)-2$. If
$T=V(H)\setminus\{u,v\}$, then $r(u|T)=(1,1,\ldots,1)=r(v|T)$.
Therefore, $T$ is not a resolving set for $G$. Since $G$ is a
randomly $k$-dimensional graph, $|T|\leq k-1$. Thus,
$\omega(G)-2=|T|\leq k-1$. Consequently, $\omega(G)\leq k+1$.
\par Clearly, if $G=K_n$, then $\omega(G)=k+1$. Conversely, let
$\omega(G)=k+1$. If $G\neq K_n$, then there exists a vertex $x\in
V(G)\setminus V(H)$ such that, $x$ is adjacent to some vertices
of $V(H)$, because $G$ is connected. Since $|V(H)|=\omega(G)$,
$x$ is not adjacent to all vertices of $V(H)$. If there exist
vertices $y,z\in V(H)$ such that, $y\nsim x$ and $z\nsim x$, then
$d(x,y)=d(x,z)=2$, because
$x$ is adjacent to some vertices of $H$. Let
$S=\{x\}\cup V(H)\setminus\{y,z\}$. Therefore,
$r(y|S)=(2,1,1,\ldots,1)=r(z|S)$. Thus, $S$ is not a resolving
set for $G$ and $|S|=k$, which is a contradiction. Hence, $x$ is
adjacent to $\omega(G)-1$ vertices of $H$.
\par On the other
hand, $x$ is adjacent to at most one vertex of $H$. Otherwise, there exist
vertices $s,t\in V(H)$ such that, $s\sim x$ and $t\sim x$. Let
$R=\{x\}\cup V(H)\setminus\{s,t\}$. Therefore,
$r(s|R)=(1,1,\ldots,1)=r(t|R)$. Thus, $R$ is not a resolving
set for $G$ and $|R|=k$, which is a contradiction. Consequently,
$\omega(G)=2$ and $k=\omega(G)-1=1$. Therefore, $G=K_2$,
which contradicts $G\neq K_n$. Hence, $G=K_n$. }\end{proof}
\begin{lemma}\label{at most k-1 common neighbor} If $res(G)=k$,
then each two vertices of $G$
have at most $k-1$ common neighbors.
\end{lemma}
\begin{proof}{ Let $u,v\in V(G)$ and $T=N(u)\cap N(v)$. Thus,
$r(u|T)=(1,1,\ldots,1)=r(v|T)$. Therefore, $T$ is not a resolving
set for $G$. Since $G$ is a randomly $k$-dimensional graph,
$|N(u)\cap N(v)|=|T|\leq k-1$.
}\end{proof}
\begin{thm}\label{delta<n-2} If $G\neq K_n$ is a randomly $k$-dimensional
graph of order $n$, then $\Delta(G)\leq n-2$.
\end{thm}
\begin{proof}{ Suppose on the contrary that there exists a vertex
$u\in V(G)$ with $deg(u)=n-1$. For each $T\subseteq
V(G)\setminus \{u\}$ with $|T|=k-1$, the set $T\cup\{u\}$ is a
resolving set for $G$ while, $T$ is not a
resolving set for $G$. Hence, there exist vertices $x,y\in
V(G)\setminus T$ such that, $r(x|T)=r(y|T)$ and $d(x,u)\neq
d(y,u)$. Since $u$ is adjacent to all vertices of $G$, we have
$u\in\{x,y\}$, say $x=u$. Thus, $r(y|T)=r(u|T)=(1,1,\ldots,1)$.
By Lemma~\ref{at most k-1 common neighbor}, $|N(u)\cap N(y)|\leq
k-1$. Hence, $deg(y)\leq k$, because $u$ is adjacent to all vertices of $G$.
This gives, $N(y)= T\cup\{u\}$.
\par Now, let
$S=T\cup\{y\}\setminus\{v\}$, for an arbitrary vertex $v\in T$.
Since $|S|=k-1$, $S$ is not a resolving set for $G$. Therefore,
there exist vertices $a,b\in V(G)\setminus S$ such that,
$r(a|S)=r(b|S)$. Since $S\cup\{u\}$ is a resolving set for $G$,
we have $d(a,u)\neq d(b,u)$. Hence, $u\in\{a,b\}$, say $b=u$.
Thus, $r(a|S)=r(u|S)=(1,1,\ldots,1)$. Consequently, $a\sim y$.
Therefore, $a\in T$, because $N(y)= T\cup\{u\}$
and $a\neq u$. Hence, $a\in(V(G)\setminus S)\cap T=\{v\}$, that is $a=v$.
Thus, $v$ is adjacent to all vertices of $T\setminus\{v\}$. Since $v$ is an
arbitrary vertex of $T$, $T$ is a clique. Therefore,
$T\cup\{u,y\}$ is a clique of size $k+1$ in $G$. Consequently, by
Theorem~\ref{omega<k+1}, $G=K_n$, which is a contradiction. Thus,
$\Delta(G)\leq n-2$.
}\end{proof}
|
\section{Introduction}
The merging processes that probably took place during the assembly of
the galaxies of the Local Group is now actively
studied \citep[e.g.][]{Klimentowski:2010}. While the Milky Way
has not suffered any major merger for several billion years and
presents only a series of dwarf galaxy stellar streams
\citep[see][]{Helmi:2008}, the other main spiral of the Local Group
(M31) has had a more perturbed history. Giant stellar loops and tidal
streams are observed in the surroundings, extending up to the
neighbour M33, which is suspected to have interacted with M31
\citep[e.g.][]{McConnachie:2009}. Many coherent structures observed
around M31 are interpreted as the disruption of small dwarf galaxies,
which are numerous in the vicinity of M31 \citep[e.g.][]{Ibata:2004}.
At the same time, it has been known for a long time that the M31 galaxy
exhibits an unusual morphology
\citep[e.g.][]{Arp:1964,Haas:1998,Helfer:2003}. The young stellar
population tracers in the disc are all concentrated in the so-called
10\,kpc ring { \citep{Chemin:2009, Nieten:2006,Azimlu:2011}}, which is
highly contrasted and superposed to only a few spiral structures {
\citep[e.g.][]{Nieten:2006,Gordon:2006}}. One striking feature is the
small amount of gas present in the central region: neither
\citet{Braun:2009} nor
\citet{Chemin:2009} detect any significant HI component, while a
similar depletion is observed in CO-emission by
\citet{Nieten:2006}. It is usually described as a quiescent galaxy
with little star formation { SFR$\sim$0.4\,\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi\,yr$^{-1}$
\citep[e.g.][]{Barmby:2006,Tabatabaei:2010,Azimlu:2011} and} with an
ultra-weak nuclear activity \citep{DelBurgo:2000}. However, ionised
gas is detected in the central field
\citep[e.g.][]{Rubin:1971,Ciardullo:1988,Boulesteix:1987,Bogdan:2008,Liu:2010},
usually interpreted in term of shocks, and \citet{Melchior:2000} has
detected only a small amount of molecular gas {($1.5\times 10^4
\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi$)} within { 1.3\,\arcmin} (305\,pc in projection).
{ There is no obvious on-going star formation in this central region
\citep[e.g.][]{Olsen:2006,Li:2009,Azimlu:2011} and the ionised gas
\citep[approximately 1500\,\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi\, according to][]{Jacoby:1985} can be
accounted for by mass lost from evolving stars.}
In this ``empty'' central region, interest has focused on the
supermassive black hole \citep{Bacon:1994}, { its activity
\citep[e.g.][]{Li:2011}} and the circumnuclear region (inner few hundred
parsecs) { \citep{Li:2009}}. \citet{Barmby:2006} observed the whole
galaxy in the mid-infrared with the {\em Infrared Array Camera} {
(IRAC)} on board the {\em Spitzer Space Telescope}, revealing
spectacular dust rings and spiral arms. \citet{Block:2006} stress the
presence of an {elongated} inner ring with projected diameters
1.5\,kpc by 1\,kpc and propose a completely new interpretation for the
morphology of this galaxy. Both rings at 1\,kpc and 10\,kpc are
off-centred. Their respective radii do not correspond to what is
expected from resonant rings in a barred spiral galaxy. The most
likely scenario for the formation of these rings is a head-on
collision of the Cartwheel type {
\citep{Struck-Marcell:1993,Horellou:2001}}. Unlike the Cartwheel,
where the companion is about 1/3rd of the mass of the target (major
merger), in Andromeda, the collision can be called a minor merger, and
produces much less contrasted rings in the main disc.
\citet{Block:2006} propose that the collision partner was M\,32, with
about 1/10th of the mass (dark matter included) at the
beginning. After stripping experienced in the collision, the M\,32
mass is now 1/23 that of the main target M\,31. The M\,32 plunging
head-on through the centre of M\,31 has triggered the propagation of
an annular wave, which is now identified with the 10\,kpc {ring}, and
a second wave propagates more slowly behind
\citep[see e.g.][]{Appleton:1996}, and would correspond to the inner
ring. In addition, the inner ring has formed {and is propagating} in
a tilted and warped disc, which accounts for its almost face-on
appearance, in contrast with the inclined main disc of M\,31. This
scenario explains why the cold gas has been expelled from the central
region and also why there are shocks and hot gas. In this article we
focus on the gas content of the inner ring and inside.
We present CO(1-0) and CO(2-1) observations obtained with the IRAM-30m
telescope in different positions located in the north-western part of
the inner ring, and a few positions inside the ring. While the CO
intensities are correlated in first instance with the A$_B$
extinction, the velocity distributions are quite unexpected. In the
inner ring the velocities are spread between -450\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi and -150\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi
for a few positions while the systemic velocity {(-310\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)} is
expected in this area along the minor axis. To better understand these
velocities, we compared our measurements with the velocities available
for the ionised {and HI} gas {and} found that they do not really
match. While neither the main HI warp of the disc {nor a bar along
the line of sight} can explain the wide amplitude velocity splittings
observed along the minor axis, the {main} configuration {compatible
with the data discussed here} is to have a ring inclined with respect
to the nuclear disc that would correspond in projection to the inner
dust ring, which is detected in extinction and in infra-red
emission. This ring could be attributed to a head-on collision with
M32 \citep{Block:2006} or to some accretion of gas from an M31-M33 gas
bridge or tidal loop detected in HI \citep{Braun:2004}.
In Section \ref{sect:obs} we describe the observations performed at
the IRAM-30m telescope. In Section \ref{sect:datared} we describe the
reduction of the C0 data. In Section \ref{sect:ana} we present an
analysis of the CO data and compare the molecular gas with the ionised
gas, together with information from other wavelengths. In Section
\ref{sect:inter} we explore various scenarios and propose one
modelling that explains the observations. In Section \ref{sect:disc} we
discuss our results and conclude.
\section{Observations}
\label{sect:obs}
\begin{figure*}
\setlength{\unitlength}{1cm}
\begin{picture}(16,11) (0,0)
\put(-0.1,5){
\begin{minipage}[l]{.2\linewidth}
\includegraphics[width=6.5cm,angle=0]{fig1a.eps}
\end{minipage} \hspace{2.5cm}
\begin{minipage}[r]{.7\linewidth}
\vspace{0.1cm}
\includegraphics[height=11.5cm,angle=-90]{fig1b.eps}
\end{minipage}
}
\put(3.15, 4.65){\line(5, 6){4.18}}
\put(7.33, 9.65){\line(5, 0){9.77}}
\put(3.25, 4.50){\line(3, -3){3.73}}
\put(6.99, 0.77){\line(5, 0){10.12}}
\end{picture}
\caption{{Right:} Positions {(J2000)} observed in CO in the
central part of M\,31. The circles indicate 1-mm (green) and 3-mm
(red) beams of the 10 positions observed superimposed on the $A_B$
extinction map obtained in \cite{Melchior:2000}. As discussed in this
paper, the extinction with the central 2 arcsec is not well-defined on
this map. The (blue) cross indicates the centre of M31
($\alpha_{J2000}$=$00^{h} 42^{m} 44^{s}.371$,
$\delta_{J2000}$=$41^{\circ} 16^{\prime} 08^{\prime\prime}.34$,
\cite{Crane:1992}.) {Left: Velocity field measured in CO for the
whole galaxy by \citet{Nieten:2006}. A small rectangle in the central
part indicates the field studied in this paper. This global view
recalls the large-scale configuration of this galaxy: a steeply
inclined disc at 77${\deg}$ with a position angle of 35${\deg}$.}}
\label{fig:super}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=6.6cm,angle=-90]{fig2a.eps}
\hspace{0.2cm}
\includegraphics[width=6.6cm,angle=-90]{fig2b.eps}
\includegraphics[width=6.6cm,angle=-90]{fig2c.eps}
\hspace{0.2cm}
\includegraphics[width=6.6cm,angle=-90]{fig2d.eps}
\includegraphics[width=6.6cm,angle=-90]{fig2e.eps}
\hspace{0.2cm}
\includegraphics[width=6.9cm,angle=-90]{fig2f.eps}
\caption{Small maps performed at in CO(2-1) (left panels) and CO(1-0)
(right panels) for the positions A, D and G. The green (resp. red)
circles correspond to the FWHM of the CO(2-1) (resp. CO(1-0)) beams at
the various positions observed. The corresponding offsets are: (0,0),
(+5,0), (+2,+4) (-2,+4), (-5,0), (-2,-4), (+2, -4). The spectra are
superimposed on the corresponding A$_B$ maps. They are positioned
arbitrarily close to the beam corresponding to the observations.}
\label{fig:sub}
\end{figure*}
The observations were carried out on 1999 June 13-15 and 2000 July
14-17 with the IRAM 30-m telescope. Most of the observations was made
in the symmetrical wobbler switching mode, in which the secondary
mirror nutates up to a maximum limit of $\pm$240 arcsec in
azimuth. The beam throw was determined as a function of the hour angle
in a way that the OFF positions lay in extinction-free regions, as
described in \cite{Melchior:2000}. It nevertheless occurred in two
positions (M31D and M31G) that some signal was detected in the wobbler
throw positions. As discussed in the Appendix \ref{sect:COOFF} and
indicated in Table \ref{tab:obs}, this corresponds to some azimuth
angles for M31G where one of the beam throws was close to the inner
spiral arm in the north/north-western part of the field. For M31D {
the configuration could be} more subtle and { the OFF signal} might
have been caused by gas located on the far side, which was not
detected in extinction. In the reduction process, the OFF signal is
smeared out in the averaging process because equal weights are taken
to reduce the observations to an equivalent beam of 24\arcsec. It is
still noticeable in M31G because the OFF signal was also present in
the 2000 observations. Near transit we had to use position-switching
mode, taking an extinction-free OFF position located at a given
position from the nucleus as indicated in the last column of Table
\ref{tab:obs}. We found the OFF signal for M31C for position-switch
observations, while no ON signal has been detected. These OFF
detections are presented in Appendix \ref{sect:COOFF}.
During the second epoch of observations small maps of seven points
with a spacing of ~5'' were made for positions (A, C, D and G) to
sample the CO(1-0) beam in CO(2-1) and to compute the CO(2-1)/CO(1-0)
line ratio. Pointing and focus calibration were regularly checked. In
total, 10 ON positions in the inner disc of M\,31 were observed in
this way, as summarised in Table \ref{tab:obs} and presented in Figure
\ref{fig:super}. Table \ref{tab:obs} provides (1) the name of the
position observed, as used throughout the article, (2) the position
angle in degrees of this position, (3) the offset in arcsec of the ON
position with respect to the centre of M\,31, (4) the distance R of
this position with respect to the centre of M\,31, (5) the right
ascension, (6) the declination of this ON position, (7) the dates when
the observations were made, (8) the total integration time on the
source $T_{exp}$, (9) if a small map (seven pointings) was made
for this position, (10) if some signal was detected in the OFF
position, if yes, WSW (resp. PSW) indicates that some OFF signal was
detected in the wobbler (resp. position) switch scans performed for
this position, (11) the offsets in arcsec used for position-switch
observations made near transit.
\begin{table*}
\caption[]{Summary of observations}
\label{tab:obs}
\begin{tabular}{p{0.05\linewidth}p{0.05\linewidth}p{0.075\linewidth}p{0.05\linewidth}p{0.075\linewidth}p{0.075\linewidth}p{0.075\linewidth}p{0.025\linewidth}p{0.025\linewidth}p{0.025\linewidth}p{0.055\linewidth}}
\hline
\noalign{\smallskip}
Position & PA & Offset &R (arcsec)$~~$ & RA (J2000) & DEC (J2000) & date & T$_{exp}$ (min) & map & Off signal & Position switch (arsec)\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
M31A & 161.1 & -26,76 & 80.3 & 00:42:42.1 & 41:17:21.1 & 13/06/99 & 630 & no & no & 122,321 \\
& & & & & & 14-17/07/00 & 1160 & yes & no & 385,57; 274,23; 379,34\\
M31C & -- & 0,0 & 0 & 00:42:44.4 & 41:16:09.2 & 2000 & 256 & yes & yes (PSW) & 385,57 \\
M31D & 161.3 & -51,151 & 159.4 & 00:42:39.9 & 41:18:39.6 & 14/06/99 & 180 & no & yes (WSW) & 122,321 \\
& & & & & & 16/07/00 & 112 & yes & no & 404,-41 \\
M31E & 47.1 & 72,67 & 98.4 & 00:42:50.8 & 41:17:15.0 & 14/06/99 & 360 & no & no & 122,321 \\
M31F & 125.6 & -134,96 & 164.8 & 00:42:32.5 & 41:17:44.0 & 15/06/99 & 442 & no & no & 122,321 \\
M31G & 121.8 & -171, 106 & 201.2 & 00:42:29.2 & 41:17:54.0 & 15/06/99 & 180 & no & yes (WSW) & 122,321 \\
& & & & & & 15/07/00 & 175 & yes & yes (WSW)& 419,-7 \\
M31GB& 170.7 & -22,134 & 135.8 & 00:42:42.4 & 41:18:22.0 & 15/06/99 & 78 & no & no & 122,321 \\
M31I & 123.8 & -172,115 & 206.9 & 00:42:29.1 & 41:18:3.6 & 17/07/00 & 42 & no & no & 540,-23 \\
M31J & 127.8 & -175,136 & 221.6 & 00:42:28.9 & 41:18:24.3 & 17/07/00 & 22 & no & no & 543,-43 \\
M31K & 143.1 & -139,185& 231.4 & 00:42:32.0 & 41:19:13.7 & 17/07/00 & 90 & no & no & 508,-93 \\
\noalign{\smallskip}
\hline
\end{tabular}
\end{table*}
We used four receivers simultaneously, two for $^{12}$CO(1-0) at
115\,GHz and two for $^{12}$CO(2-1) at 230\,GHz. At 115\,GHz each
receiver was connected to two autocorrelator sub-bands (shifted by
40\,MHz from each other) and each sub-band consisted of 225 channels
separated by 1.25\,MHz. At 230\,GHz each receiver was connected to a
filter-bank consisting of 512 channels of 1\,MHz width.
\begin{table*}
\caption[]{Characteristics of the CO lines. CO(2-1)
(resp. CO(2-1)) spectra are reduced to a 2.6\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi
(resp. 3.2\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi) resolution.}
\label{tab:lines}
\begin{tabular}{p{0.1\linewidth}p{0.1\linewidth}p{0.12\linewidth}p{0.1\linewidth}p{0.1\linewidth}p{0.05\linewidth}p{0.05\linewidth}p{0.1\linewidth}p{0.05\linewidth}}
\hline
\noalign{\smallskip}
Position & $^{12}$CO line (beam size) & I$_{CO}$ (K km s$^{-1}$) $=\int T_{mb} dV$& V$_0$ (km s$^{-1}$) & $\sigma$ (km s$^{-1}$) & T$_{\rm peak}$ (mK) & baseline rms (mK) & $N_{H_2}$ (cm$^{-2}$) & $\Sigma_{H_2}$ (M$_\odot$\,pc$^{-2}$)\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
M31A & 1-0 (24$\arcsec$) & 0.75$\pm$0.03 & -154.9$\pm$0.6 & 28.7$\pm$1.3 & 24.8 & 2.1& 1.73$\times 10^{20}$ & 2.94\\
& 2-1 (24$\arcsec$) & 0.88$\pm$0.06 & -151.1$\pm$0.7 & 22.9$\pm$1.9 & 35.7 & 4.7& $-$ & \\
& 1-0 (21$\arcsec$) & 0.72$\pm$0.04 & -153.6$\pm$0.7 & 28.7$\pm$1.7 & 23.7 & 2.4& 1.67$\times 10^{20}$ & 2.83\\
& 2-1 (11$\arcsec$) & 0.78$\pm$0.06 & -152.8$\pm$1.0 & 24.9$\pm$2.6 & 29.4 & 5.5& $-$ & \\
M31C & 1-0 (24$\arcsec$) & $-$ & $-$ & $-$ & $-$ & 3.7& $<$2.55$\times 10^{21}$& $<$43.31\\
& 2-1 (24$\arcsec$) & $-$ & $-$ & $-$ & $-$ & 7.8& $-$ & \\
M31D & 1-0 (24$\arcsec$) & 2.43$\pm$0.10 & -74.3$\pm$0.4 & 23.7$\pm$1.2 & 96.4 & 7.2& 5.58$\times 10^{20}$ & 9.49\\
& 2-1 (24$\arcsec$) & 3.45$\pm$0.16 & -73.2$\pm$0.5 & 19.7$\pm$1.1 & 164.6 &15.5& $-$ & \\
& 1-0 (21$\arcsec$) & 2.47$\pm$0.11 & -74.9$\pm$0.5 & 23.6$\pm$1.4 & 98.0 & 8.7& 5.68$\times 10^{20}$ & 9.66\\
& 2-1 (11$\arcsec$) & 2.12$\pm$0.12 & -74.9$\pm$0.6 & 21.1$\pm$1.4 & 93.8 &10.6& $-$ & \\
M31E$^{\mathrm{a}}$ & 1-0 (21$\arcsec$) & 0.18$\pm$0.04 & -132.2$\pm$1.9 & 15.8$\pm$3.1 & 10.5 & 4.1& 4.25$\times 10^{19}$ & 0.72\\
& 2-1 (11$\arcsec$)& 0.33$\pm$0.10 & -128.2$\pm$3.1 & 15.3$\pm$6.9 & 19.4 &13.1& $-$ & \\
M31F & 1-0 (21$\arcsec$) & 1.56$\pm$0.14 & -360.6$\pm$3.3 & 68.0$\pm$6.7 & 20.6 & 6.4& 3.59$\times 10^{20}$ & 6.11\\
& 2-1 (11$\arcsec$) & $-$ & $-$ & $-$ & $-$ &16.3& $-$ & \\
M31G & 1-0 (24$\arcsec$) & 2.98$\pm$0.18 & -147.4$\pm$0.4 & 13.6$\pm$1.0 & 205.9 &4.9& 6.86$\times 10^{20}$ & 11.66\\
& 2-1 (24$\arcsec$) & 2.39$\pm$0.14 & -146.7$\pm$0.5 & 15.4$\pm$1.1 & 145.2 &9.8& $-$ & \\
& 1-0 (21$\arcsec$) & 3.01$\pm$0.23 & -146.7$\pm$0.6 & 15.3$\pm$1.4 & 185.2 &5.3& 6.92$\times 10^{20}$ & 11.77\\
& 2-1 (11$\arcsec$) & 2.18$\pm$0.20 & -146.6$\pm$0.9 & 18.5$\pm$1.9 & 110.6 &15.9& $-$ & \\
M31GB & 1-0 (21$\arcsec$) & $-$ & $-$ & $-$ & $-$ & 7.7& $<$5.29$\times 10^{21}$& $<$89.95\\
& 2-1 (11$\arcsec$) & $-$ & $-$ & $-$ & $-$ &14.7& $-$ & \\
M31I & 1-0 (21$\arcsec$) & 3.22$\pm$0.11 & -148.5$\pm$0.2 & 9.3$\pm$0.4 & 324.3 &12.5& 7.41$\times 10^{20}$ & 12.60\\
& 1-0 (21$\arcsec$) & 4.60$\pm$0.77 & -411.9$\pm$1.0 & 34.1$\pm$3.1 & 126.8 &12.5& 1.06$\times 10^{21}$ & 17.99\\
& 1-0 (21$\arcsec$) & 2.50$\pm$0.31 & -278.7$\pm$2.6 & 46.3$\pm$5.4 & 50.7 &12.5& 5.75$\times 10^{20}$ & 9.77\\
& 1-0 (21$\arcsec$) & 3.69$\pm$0.89 & -376.4$\pm$11.1& 88.5$\pm$16.5& 39.2 &12.5& 8.49$\times 10^{20}$ & 14.44\\
& 2-1 (11$\arcsec$) & 3.28$\pm$0.12 & -148.6$\pm$0.2 & 7.7$\pm$0.4 & 400.9 &29.0& $-$ & \\
& 2-1 (11$\arcsec$) & 7.28$\pm$0.47 & -409.5$\pm$1.3 & 40.0$\pm$3.5 & 171.0 &29.0& $-$ & \\
& 2-1 (11$\arcsec$) & 4.04$\pm$0.47 & -283.4$\pm$2.7 & 43.5$\pm$5.0 & 87.7 &29.0& $-$ & \\
& 2-1 (11$\arcsec$) & 2.96$\pm$0.41 & -353.3$\pm$5.7 & 46.6$\pm$4.5 & 59.6 &29.0& $-$ & \\
M31J & 1-0 (21$\arcsec$) &12.81$\pm$0.33 & -413.2$\pm$0.2 & 33.3$\pm$0.8 & 361.4 &12.2& 2.95$\times 10^{21}$ & 50.08\\
& 1-0 (21$\arcsec$) & 1.06$\pm$0.11 & -145.5$\pm$0.6 & 11.9$\pm$1.5 & 83.8 &12.2& 2.45$\times 10^{20}$ & 4.16\\
& 1-0 (21$\arcsec$) & 7.34$\pm$0.57 & -350.1$\pm$7.5 &178.2$\pm$13.2& 38.8 &12.2& 1.69$\times 10^{21}$ & 28.70\\
& 2-1 (11$\arcsec$) &15.22$\pm$0.78 & -412.7$\pm$0.9 & 35.1$\pm$2.0 & 407.0 &54.5& $-$ & \\
& 2-1 (11$\arcsec$) & 0.86$\pm$0.41 & -142.1$\pm$3.0 & 9.0$\pm$5.2 & 88.5 &54.5& $-$ & \\
& 2-1 (11$\arcsec$) & 1.94$\pm$0.69 & -268.6$\pm$7.1 & 35.4$\pm$10.8& 51.4 &54.5& $-$ & \\
M31K & 1-0 (21$\arcsec$) & 2.66$\pm$0.14 & -300.8$\pm$1.1 & 41.7$\pm$2.9 & 59.8 & 7.5& 6.11$\times 10^{20}$ & 10.38\\
& 2-1 (11$\arcsec$) & 2.61$\pm$0.43 & -292.5$\pm$4.9 & 58.5$\pm$11.0& 42.0 &24.1& $-$ & \\
\noalign{\smallskip}
\hline
\end{tabular}
\begin{list}{}{}
\item[$^{\mathrm{a}}$] Tentative detection.
\end{list}
\end{table*}
\section{Data reduction}
\label{sect:datared}
The {\sc{CLASS}}\footnote{Continuum and Line Analysis Single-dis
Software, http://www.iram.fr/IRAMFR/GILDAS} package was used for
the data reduction. After checking the quality of each single spectra,
the data were averaged with inverse variance weights. A
first-order baseline was fitted to the resulting spectrum and
subtracted. For a few spectra, a higher-order polynomial was
carefully subtracted. Finally, the spectra were smoothed to a
velocity resolution of 3.2\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi (resp. 2.6\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi) in
CO(1-0) (resp. CO(2-1)).
The maps are presented in Figure \ref{fig:sub} and illustrate the
observing procedure. The spectra are located arbitrarily close to the
beam position they correspond to. Spatial variations in the spectra
agree relatively well with the extinction distribution. For M31G the
CO intensity is clearly stronger in the area where the extinction is
stronger. For M31A and M31D the intensity of the lines does not follow
the intensity of the extinction exactly but we cannot exclude that
some pointing errors and/or a clumpy distribution of the underlying
gas explains the spatial variations of the CO distribution.
The results of the fitting are presented in Table \ref{tab:lines}. A
Gaussian function is fitted to each line to determine its area,
central velocity $V_0$, width $\sigma$, and peak temperature $T_{\rm
peak}$. The baseline RMS is provided for each line. We provide
main-beam temperatures (unless specified otherwise) throughout this
paper with B$_{eff}=$ 64.2$\pm 3$ and F$_{eff}=$ 91$\pm 2$
(resp. B$_{eff}=$ 42.$\pm 3$ and F$_{eff}=$ 86$\pm 2$) at 115\,GHz
(resp. 230\,GHz). For the positions for which maps have been observed,
the spectra were convolved to obtain an 24$\arcsec$ beam ($\sqrt{11^2
+ 21^2}$) for both lines, enabling a direct computation of the line
ratio.
The spectra are displayed in Figure \ref{fig:M31all}. In the final
spectra, the signal present in the OFF positions is mainly visible for
the position M31G. It is at the systemic velocity and probably
corresponds to the gas in the main disc along the minor axis in the
north of the M31G position (because the OFF positions are symmetric in
azimuth with respect to the ON position). { We use probably throughout
this paper except for the \citet{Saglia:2010} data, whose velocities
are formally in LSR, but the difference ($\sim$5km\,s$^{-1}$) is
negligible here.}
\begin{figure*}
\centering
\includegraphics[width=6.3cm,angle=-90]{fig3a.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3b.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3c.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3d.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3e.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3f.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3g.eps}
\includegraphics[width=6.3cm,angle=-90]{fig3h.eps} \caption{Spectra
with a detected signal. M31A, M31D and M31G (resp. M31E, M31F, M31I
and M31J) spectra are reduced to a 24$\arcsec$ beam (resp. with the
original beam) in CO(2-1) and CO(1-0). The y-axis provides the
antenna temperature in K.} \label{fig:M31all}
\end{figure*}
\section{Analysis}
\label{sect:ana}
\begin{table*}
\caption[]{A$_B$ extinction derived from the map based on the
\citet{Ciardullo:1988} data as obtained in
\citet{Melchior:2000}. For each configuration E(B-V) was then
computed assuming R$_V$ in the range $[2.1,3]$. These
values were computed assuming a fraction of foreground
light $x=0$. The I$_{CO}/E(B-V)$ ratio is then provided for
the strongest lines. For M31J we also provide the ratio
corresponding to the ring velocity.}
\label{tab:abs}
\begin{tabular}{llllllll}
\hline
\noalign{\smallskip}
Position & A$_B$ at 24$\arcsec$ & A$_B$ at 21$\arcsec$ (11$\arcsec$) &
E(B-V) at 24$\arcsec$ & E(B-V) at 21$\arcsec$ & E(B-V) at
11$\arcsec$ & I$_{CO}$/E(B-V) at 24$\arcsec$ &I$_{CO}$/E(B-V) at 21$\arcsec$ \\
& (mag beam$^{-1}$) & (mag beam$^{-1}$) & (mag
beam$^{-1}$) & (mag beam$^{-1}$) & (mag beam$^{-1}$) & (K km s$^{-1}$
mag$^{-1}$)& (K km s$^{-1}$
mag$^{-1}$)\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
M31A & 0.095 & 0.100 (0.166) & 0.027$\pm$0.003 &0.029$\pm$0.004 &0.048$\pm$0.006& 28$\pm$5 & 26$\pm$ 5\\
M31D & 0.067 & 0.076 (0.111) & 0.019$\pm$0.002 &0.022$\pm$0.003 &0.032$\pm$0.004&129$\pm$22&116$\pm$20\\
M31F & 0.074 & 0.081 (0.123) & 0.021$\pm$0.003 &0.023$\pm$0.003 &0.035$\pm$0.004& & 69$\pm$15\\
M31G & 0.182 & 0.193 (0.256) & 0.052$\pm$0.007 &0.055$\pm$0.007 &0.073$\pm$0.009& 59$\pm$11& 56$\pm$11\\
M31I & 0.210 & 0.219 (0.245) & 0.060$\pm$0.008 &0.063$\pm$0.008 &0.070$\pm$0.009& & 52$\pm$ 8\\
M31J & 0.203 & 0.208 (0.241) & 0.058$\pm$0.007 &0.060$\pm$0.008 &0.069$\pm$0.009& & 219$\pm$33(18$\pm$ 4)\\
M31K & 0.138 & 0.141 (0.159) & 0.040$\pm$0.005 &0.040$\pm$0.005 &0.046$\pm$0.006& & 67$\pm$12\\
\noalign{\smallskip}
\hline
\end{tabular}
\end{table*}
\subsection{Characteristics of the CO emission and gas-dust connection}
The CO gas, which we searched for and discuss in this paper, is
located mainly around the minor axis of the main disc of M31. The
detected molecular gas complexes correspond to the strongest A$_B$
extinction complexes that we observed. For the positions without
extinction (M31GB), we detected no signal, while no
line was detected in the centre either. For M31E where the
extinction is weak, we have a tentative detection at 4$\sigma$. The
extinction wascomputed assuming a fraction of foreground light
of $x=0$ \citep[see][]{Melchior:2000}, which is probably correct for
the north-western part, where we observed, because an inclination of
45\,$\deg$ is usually assumed for the nuclear spiral
\citep{Ciardullo:1988}. {We also computed the near-UV and far-UV
extinction maps from archive GALEX data with the same method and found
the same structures. In addition, we subtracted from the original
GALEX images the bulge emission modelled in this way, and found no
trace of UV emission in the inner ring.}
\begin{figure*} \centering
\includegraphics[height=14cm,angle=-90]{fig4.eps}
\caption{Superimposition of the G, I, and J spectra at the nominal
spatial resolution, namely 21$\arcsec$ for CO(1-0) and 11$\arcsec$
for CO(2-1). The y-axis provides the
antenna temperature in K.} \label{fig:superimp} \end{figure*}
The position M31F, which exhibits a weak CO(1-0) signal with no
detection in CO(2-1), corresponds to a position with low A$_B$
extinction. The other positions are detected with high signal-to-noise
ratios at least in CO(1-0). We computed the A$_B$ extinction values
for each position convolved with the different beams, as provided in
Table \ref{tab:abs}, and there is no one-to-one correspondence with
the intensity of the detected CO. Figure \ref{fig:super} also shows
that M31D has a much stronger signal in CO than M31A while its A$_B$
extinction is weaker. One plausible explanation could be that the dust
clumps corresponding to M31D do not lie in the same plane as M31A: if
the foreground light ($x$) exceeds 0, we have underestimated the
extinction. Assuming the I$_{CO}$/A$_B$ ratio measured for M31A
applies to M31D, one would expect a peak intensity A$_B^{real} = 1.56$
for M31D, but it is measured A$_B^{meas.} = 0.26$. This configuration
corresponds to a fraction of light in front of the dust $x=0.72$,
meaning that M31D lies on the back side of the bulge. It is also
probable that the gas is very clumpy and that the non-linearity of the
extinction somehow biases our A$_B$ estimate.
Our detections are concentrated in an area\footnote{As in
\citet{Melchior:2000}, we assume a distance of M\,31 of 780\,kpc
i.e. 1\,arcsec$=$3.8\,pc.} of 415\,pc $\times$ 570\,pc (in
projection), located in the north-western part of the M\,31 within
3.8\,arcmin from the centre, corresponding to 880\,pc (resp. 1.2\,kpc
if deprojected {for a 45$\deg$ inclination}). While the area explored
is quite localised, the detected velocities span from -73 to
-413\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, while velocities along the minor axis of a rotating disc
are expected to be at the systemic velocity. M31K is the only position
exhibiting this expected velocity, as well as a component for M31G,
M31I and M31J, and possibly M31D. Figure \ref{fig:superimp} shows a
superimposition of the spectra obtained for M31G, M31I and M31J (and
reduced to a 24\arcsec\, beam): the various velocity patterns appear
in each spectra with different relative intensities. Besides the
velocity line detected at -145\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, a broadband signal is detected
with velocities between -450 and -250\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. As discussed below, this
range of velocities cannot be explained by a single disc inclined at
45\,deg in regular rotation, and they are the signature of a peculiar
structure. The line widths of the detected lines are quite different
from one location to another, suggesting a spatial extension of the
emitting areas. We assume a Galactic $X_{CO} = N_{H_2}/I_{CO} = 2.3
\times 10^{20}$\,cm$^{-2}$(K\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)$^{-1}$ following
\citet{Strong:1988}. For the positions where we did not detect any
signal, we provide a 3\,$\sigma$ upper limit based on the
dispersion (rms) computed on the baseline. Some CO(2-1) measurements
(M31I, M31J, M31K) are at the limit of detection and are provided for
a rough comparison with CO(1-0) measurements.
In the centre {($<2\arcsec$)}, our $A_B$ map does not measure the
extinction (owing to the method, which is based on ellipse fitting).
The map of 8\,$\mu$m emission (after subtraction of the scaled stellar
continuum at 3.6\,$\mu$m) also displays a defect in the central part,
so it is difficult to be conclusive about the extinction in the inner
arcseconds. However, \citet{Garcia:2000}, relying on simple modelling
of {\em Chandra} data, have estimated $A_V=1.5\pm 0.6$ in the central
3-arcsec region. Our millimetre observations exclude the presence of
CO(1-0) (resp. CO(2-1)) at the 1\,$\sigma$ level of 3.7\,mK
(resp. 7.8\,mK). On the basis of this upper limit, the gas present
within 80\,pc from M31's centre does not exceed
43\,M$_\odot$/pc$^{-2}$ (3\,$\sigma$), assuming a conservative value
for X$_{CO}$ typical of the Galactic disc. This is an extremely small
amount compared to the gas mass computed in the central region of the
Milky Way \citep{Oka:1998}.
\begin{table}
\caption[]{Line ratios of the complexes reduced to a 24$\arcsec$
beam. We provide both the ratios of the line intensities r$_{12}$ and the
ratios of the peak temperatures $R$.}
\label{tab:linerat}
\begin{tabular}{lll}
\hline
\noalign{\smallskip}
Complexes & r$_{12}$ & $R$ \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
M31A & 1.17$\pm^{0.40}_{0.19}$ & 1.44$\pm^{0.3}_{0.1}$ \\
M31D & 1.43$\pm^{0.45}_{0.21}$ & 1.71$\pm^{0.36}_{0.12}$ \\
M31G & 0.80$\pm^{0.30}_{0.13}$ & 0.71$\pm^{0.14}_{0.05}$ \\
\noalign{\smallskip}
\hline
\end{tabular}
\end{table}
We used the three positions with observations reduced to a 24\arcsec\,
beam to compute the CO(2-1) to CO(1-0) line ratio. As displayed in
Table \ref{tab:linerat}, we computed both the ratio of the integrated
intensities and the ratio of the peak temperatures, which are
compatible within error bars. M31G exhibits a line ratio comparable
with the mean value found for the Milky Way spiral arms
\citep{Sakamoto:1997}.
M31A and M31D exhibits CO line ratios higher than 1. Given the
N$_{H_2}$ column densities computed (respectively $5.6\times 10^{20}$
cm$^{-2}$ and $1.7\times 10^{20}$ cm$^{-2}$), the gas is probably not
optically thin. In addition, the velocity dispersion of CO(2-1) is
systematically smaller than that of CO(1-0), suggesting that these
complexes are probably externally heated and composed of dense
gas. This hypothesis is compatible, as discussed below, with the
combined presence of an intense stellar radiation field caused by
bulge stars \citep{Stephens:2003}, the outflow detected in X-ray in
this area \citep{Bogdan:2008}, and the ionised gas
\citep{Jacoby:1985}. This supports the view that no star formation is
currently happening inside the inner ring of M31.
\begin{figure*}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig5a.eps}
\hspace{0.2cm}
\includegraphics[width=6.5cm,angle=-90]{fig5b.eps}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig5c.eps}
\hspace{0.2cm}
\includegraphics[width=6.5cm,angle=-90]{fig5d.eps}
\caption{H$\alpha$+[NII] (left) and [NII] (right) intensities from
\citet{Ciardullo:1988} and \citet{Boulesteix:1987} are displayed
with an arbitrary normalisation in the top panels. The bottom left
panel presents the CO velocities measured in this paper,
superimposed on the gas kinematics based on (1) H$\beta$ and
$[$O{\sc iii}$]$ lines measured in slits by \citet{Saglia:2010},
(2) $[$O{\sc ii}$]$ and $[$Ne{\sc iii}$]$ measured in slits by
\citet{Ciardullo:1988} and (3) H$\alpha$ and $[$N{\sc ii}$]$
measured in slits by \citet{Rubin:1971}. The bottom right panel
displays the velocity field obtained with $[$N{\sc ii}$]$ $\lambda
6584\dot{A}$ observations (cf top right panel) based on a
Fabry-Perot device by \citet{Boulesteix:1987}. On each panel, the
positions of our CO observations are indicated with circles
corresponding to the beams.} \label{fig:HaNIIvel}
\end{figure*}
\subsection{Ionised gas}
The ionised gas in the central regions of M\,31 has been mapped by
\citet{Ciardullo:1988}, who produced the highest S/N ratio map of
H$\alpha$+[NII] based on the compilation of a five-year nova survey
published so far. It exhibits, as displayed in the top left panel of
Figure \ref{fig:HaNIIvel}, a turbulent spiral, which appears more
face-on than does the main disc. These authors described it as lying
in a disc, warped in a way that the gas south of the nucleus is viewed
closer to face-on than the gas in the northern half of the galaxy. As
discussed in \citet{Block:2006}, this arc feature corresponds to the
south-west part of the offset ring detected with {\em Spitzer}{-IRAC}
data. \citet{Rubin:1971} measured that in this area [NII] is three
times stronger than H$\alpha$, which supports the shock hypothesis as
also discussed by \citet{Jacoby:1985}. According to \citet{Liu:2010},
half of the kinematic temperature of hot gas in the central bulge is
accounted for by the stellar dispersion, but additional heating is
expected from type-Ia supernovae, even though the question of the iron
enrichment is not entirely clear. A head-on collision with M32
suggested by \citet{Block:2006} could account for the additional
heating and at least contribute to it.
In parallel, \citet{Boulesteix:1987} observed this same area with a
Fabry-Perot spectrograph to map the velocity field, and produced an
[NII] map, which is xdisplayed in the top right panel of Figure
\ref{fig:HaNIIvel}. The centre has been masked for technical reasons.
Besides the lower resolution and lower S/N ratio than for the
\citet{Ciardullo:1988} map (because of a smaller integration time), it is
instructive to observe that it exhibits different patterns compared
with the H$\alpha$+[NII] map. It is striking that the filament, which
crosses the position M31A, is not clearly detected in [NII], so it
should be mainly associated to H$\alpha$ component. The comparison of
these two maps enhances the fact that obviously forbidden lines and
Balmer lines do not sample exactly the same regions: (1) the gas does
not have the same H$\alpha$/[NII] ratio { everywhere}; (2) the
extinction additionally complicates the analysis.
\subsubsection{Velocity field of the central bulge (1.5\,kpc$\times$1.5\,kpc)}
To compare the ionised gas with our molecular detections, we tried to
reconstruct the ionised gas velocity field from data published in the
literature. Various lines have been used, but only a single velocity
has been measured for each spectrum. In the bottom right panel of
Figure \ref{fig:HaNIIvel} we display the velocity map of
\citet{Boulesteix:1987}. The best S/N ratio is of course observed
where the [NII] intensity is strongest (cf. top right panel of
Fig. \ref{fig:HaNIIvel}). It exhibits an irregular disc in rotation,
with obvious perturbations along the minor axis. Planetary nebulae
are numerous in this field \citep[537 detected by][]{Merrett:2006} and
add noise to this velocity field. As discussed below (see Table
\ref{geometry}), we estimate here a position angle of 70$\deg$
(-20$\deg$ for the kinematic minor axis) compared to 40$\deg$ claimed
by \citet{Boulesteix:1987}. \citet{Saglia:2010} interpreted these
perturbations as counter-rotations. In the bottom left panel of Figure
\ref{fig:HaNIIvel}, we display the slit measurements we gathered from
the literature, using
Dexter\footnote{http://dc.zah.uni-heidelberg.de/sdexter} when
necessary, namely H$\beta$ and $[$O{\sc iii}$]$ from
\citet{Saglia:2010}; $[$O{\sc ii}$]$ and $[$Ne{\sc iii}$]$ from
\citet{Ciardullo:1988}; H$\alpha$ and $[$N{\sc ii}$]$ from
\citet{Rubin:1971}. We also superimpose our CO velocities, and find a very
marginal agreement. In Figure \ref{fig:XHaboul} we superimpose the
two figures. We find a good overall agreement for the various ionised
gas measurements with complicated features in the central part and
along the minor axis, with the most striking discrepancy at position
angles of 128\,$\deg$ and 142\,$\deg$.
It is difficult to understand the discrepancies observed in the ionised
gas velocity field, as the techniques are quite different and the
\citet{Boulesteix:1987} data suffers relatively low resolution (even
though excellent given the fact that it has been obtained in 1985!).
We can consider several explanations: (1) the various lines might have
different relative ratios from one region to another (as seen in
the top panels of Figure \ref{fig:HaNIIvel}) and might be affected
differently by extinction (see Figure \ref{fig:super}). (2) It
is possible that the ionised gas, like the molecular gas, exhibits
several velocity components, and it has been not explored so far, but
by \citet{Boulesteix:1987}, who mentioned line splittings greater than
30\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi in the central region.
In addition, the CO velocities do not follow the regular pattern. M31A
and M31D do not really match the ionised gas velocities. However, some
CO positions might have a velocity (or at least one component)
compatible with the ionised gas, namely M31E, M31F, M31K, M31I, M31J
and M31G. This suggests that only part of the molecular gas is
kinetically decoupled from the ionised gas.
Also, we wanted to figure out whether similar line splittings were
present in the ionised data. As discussed by \citet{Saglia:2010}, the
intrinsic velocity dispersions of the gas is smaller than 80\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi,
while the instrumental resolution achieved by these authors is
57\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. \citet{Boulesteix:1987}, who detected line splittings larger
than 30\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\, in the central area, reached a spectral resolution of
14\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. It is thus challenging to detect line splittings of this
order of magnitude, but such splittings might be underlying and
explain part of the discrepancies. As we observe large ($>200$\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)
line splittings in CO, we try to investigate if such double components
can be detected from the ionised gas.
\begin{table*}
\caption[]{Line splittings detected on $[$O{\sc iii}$]$ and H$\beta$
lines from \citet{Saglia:2010}. We provide the velocities,
signal-to-noise ratio of the integrated line and $[$O{\sc iii}$]$/H$\beta$ line
ratio {($\mathbf{\log_{10}}$)} for each detected component (1 and 2). We also give for
comparison purposes the single values published by \citet{Saglia:2010}.}
\label{tab:linespl}
\begin{tabular}{crr|crc|crc|c}
\hline
\noalign{\smallskip}
offsets & R($\arcsec$) & PA($\deg$) &
v$_1^{[\mathrm{O{\sc iii}}]5007}$ (\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi) & S/N$_1$ & $[$O{\sc iii}$]$/H$\beta$
& v$_2^{[\mathrm{O{\sc iii}}]5007}$ (\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi) & S/N$_2$ &
$[$O{\sc iii}$]$/H$\beta$ & v$^{\mathrm{gas}}_{\mathrm{Saglia}}$ \\
\noalign{\smallskip}
\hline
-15.2,16.9 & -22.8& 138 &-548.6$\pm$12.0& 4.7 & 0.51$^{+0.20}_{-0.19}$ & -360.6$\pm$16.7& 5.2&-0.22$^{+0.10}_{-0.12}$&-469.1$\pm$8.4\\
-13.5,-12.2& -18.2& 48 &-462.8$\pm$6.0 & 13.1 & 0.29$^{+0.09}_{-0.08}$ & -234.0$\pm$4.0 &15.4& 0.79$^{+0.37}_{0.22}$ &-323.7$\pm$7.4\\
-12.0,-10.8& -16.2& 48 &-449.5$\pm$9.1 & 10.4 & 0.73$^{+0.20}_{-0.16}$ & -225.4$\pm$5.2 &12.2& 0.86$^{+0.32}_{0.21}$ &-314.8$\pm$5.4\\
-9.3,-8.4 & -12.5& 48 &-485.8$\pm$9.9 & 4.9 & 1.17$^{+\inf}_{-0.71}$ & -375.0$\pm$6.7 & 3.9&-0.23$^{+0.20}_{0.21}$ &-418.5$\pm$10.7\\
-8.1,-7.3 & -10.9& 48 &-454.3$\pm$14.4& 3.3 & 0.15$^{+0.23}_{-0.25}$ & -272.5$\pm$13.3& 5.3& 0.41$^{+0.22}_{0.20}$ &-314.4$\pm$9.2\\
-1.9,-5.9 & -6.2 & 18 &-319.9$\pm$20.0& 3.3 & 0.14$^{+0.34}_{-0.31}$ & -144.8$\pm$15.9& 3.6 & 0.19$^{+0.32}_{0.28}$&-342.7\\
-0.2,0.9 & -0.9 & 168 &-281.3$\pm$8.9 & 8.7 & 0.78$^{+1.05}_{-0.33}$ & -85.9$\pm$16.2 & 3.2 & 0.00$^{+0.72}_{0.41}$&-241.8$\pm$10.0\\
0.1,-0.7 & 0.7 & 168 &-319.9$\pm$19.8& 3.3 &-0.20$^{+0.14}_{-0.18}$ & -144.8$\pm$16.1& 3.7 &-0.04$^{+0.20}_{0.22}$&-221.7$\pm$11.6\\
0.7,0.6 & 1.0 & 48 &-332.6$\pm$11.0& 6.8 & 0.60$^{+0.22}_{-0.18}$ & -64.9$\pm$7.5 & 7.5& 1.04$^{+\inf}_{-0.40}$&-313.6$\pm$7.8\\
1.1,1.0 & 1.5 & 48 &-322.3$\pm$13.0& 4.9 &-0.06$^{+0.17}_{-0.17}$ & -56.0$\pm$3.5 &18.2& 0.92$^{+0.32}_{0.20}$ &-330.6$\pm$7.8\\
1.2,0.3 & 1.2 & 78 &-321.5$\pm$13.2& 5.2 & 0.20$^{+0.26}_{-0.22}$ & -78.7$\pm$4.8 &12.7& 0.57$^{+0.22}_{0.17}$ &-327.8$\pm$11.9\\
8.8,-2.9 & 9.3 & 108 &-598.0$\pm$4.7 & 14.1 &0.92$^{+0.37}_{-0.22}$ & -319.2$\pm$14.0& 8.1 & 0.18$^{+0.12}_{-0.12}$&-343.6$\pm$13.9\\
11.2,-3.6 & 11.7 & 108 &-593.7$\pm$5.5 & 8.4 &1.17$^{+0.56}_{-0.28}$ & -373.3$\pm$29.2& 3.7 &-0.18$^{+0.16}_{-0.19}$&-376.8$\pm$18.1\\
5.3,16.4 & 17.2 & 18 &-575.1$\pm$3.8 & 6.4 &-0.16$^{+0.17}_{-0.16}$ & -444.8$\pm$35.2& 4.5 &0.28$^{+0.17}_{-0.18}$&-291.4$\pm$6.6\\
\noalign{\smallskip}
\noalign{\smallskip}
\hline
\end{tabular}
\end{table*}
\begin{figure}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig6.eps}
\caption{6.7$^{'}\times$6.7$^{'}$ ionised gas velocity field
centred on M31. Superimposition of the various gas velocities (see
Fig. \ref{fig:HaNIIvel} bottom left) on the \citet{Boulesteix:1987}
Fabry-Perot-derived velocity map.} \label{fig:XHaboul}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig7.eps}
\caption{40$^{''}\times$40$^{''}$ ionised gas velocity field
centred on M31. Line splittings detected in $[$O{\sc
iii}$]$\,50007$\dot{A}$ are superimposed on the velocity field
measured on ionised gas by \citet{Saglia:2010} and by
\citet{DelBurgo:2000}. This velocity field is quite unexpected and
do not display any clear rotation pattern. } \label{fig:HLS}
\end{figure}
Optical spectra of M\,31 are heavily dominated by the stellar
continuum of its bulge. Therefore, in order to study emission lines in
that spectral region we approximated the calibrated sky subtracted
spectra of M\,31 kindly provided by R.~Saglia by stellar population
models using the {\sc NBursts} full spectral fitting technique
\citep{CPSA07,CPSK07}. We excluded narrow regions around H$\beta$,
$[$O{\sc iii}$]$, and $[$N{\sc i}$]$ lines from the fitting. Then we
subtracted the best-fitting stellar population models from the
original spectra and analysed the fitting residuals. This allowed us
to reliably measure parameters of these rather faint emission lines
which are barely visible in the original data. We then inspect
visually the spectra to determine the positions where high S/N line
splittings is present. We then fit a two Gaussian functions with the
CLASS package on the $[$O{\sc iii}$]$5007$\dot{A}$ line. Template mismatch and
the relative low S/N affect the measurement of the H$\beta$, which is
weaker and not always detected. For none of our measurements (but
one), the [NI] line is detected. Our detection are summarised in Table
\ref{tab:linespl} and Figure \ref{fig:HLS}. We identify 14 positions
with a double component detected in $[$O{\sc iii}$]$ and also in
H$\beta$, within 23\arcsec\, ($\sim$86\,pc) from the centre, but we do not
detect any line splittings close to our CO observations or at a
similar radial distance.
\subsubsection{ Velocity field of the circumnuclear region
(150\,pc$\times$150\,pc)}
\label{sssect:linesplit}
Figure \ref{fig:HLS} displays the velocity field measured in
ionised gas of the circumnuclear region. This includes slit data from
\citet{Saglia:2010} as well as two-dimensional spectroscopy of
\citet{DelBurgo:2000} for the measurements of single velocity
components, and our line splitting measurements based on
\citet{Saglia:2010} data. It is striking that
unlike Figure \ref{fig:XHaboul} this velocity field does not exhibit
any clear rotation pattern in the central field. While the centre
presents a spot at the systemic velocity, the whole area has
velocities smaller than the systemic velocity and is approaching the
observer. This coherent flow of ionised gas is decoupled from the
stellar kinematics
\citep[e.g.][]{Bender:2005,DelBurgo:2000,Saglia:2010}, and resembles
an ionised gas outflow. This is confirmed by the fact that the two
independent data sets are in good agreement. { We should stress
that the \citet{Saglia:2010} data are characterised by an instrumental
resolution of 57\,km\,s$^{-1}$, and their systemic velocity is shifted
by $+23$\,km\,s$^{-1}$ with respect to the other studies. However, we
checked that this cannot account for the systematic blue shift
observed with respect to the centre.} The mechanisms that rule the
feed-back from the AGN are still largely unknown and in the case of
M31, we are only tracing a relic activity of its known black
hole. Considering that M\,31's black hole is not active, two main
mechanisms can be considered to explain this outflow. (1) It could be
directly associated to the past stellar activity detected in the
central regions: a 200Myr old stellar disc with a 2400$\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi$ mass has
been detected by \citet{Bender:2005} within a fraction of arcsec from
the centre. Inside a 2\arcsec disc,
\citet{Saglia:2010} has also detected a stellar component younger than
600\,Myr corresponding to a mass smaller than 2$\times 10^{6}$\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi:
it could be compatible with a 100\,Myr component with a mass of
$10^{6}$\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi. However, {we lack of constraints to conclude for} a
star formation origin {with a typical} dynamic time $\tau_{dyn} = $
size $/$ velocity of this outflow. {The typical velocity of an outflow
\citep[e.g.][]{Rupke:2005} is about 400\,km\,s$^{-1}$, a
$\tau_{dyn}=100$\,Myr old starburst would then correspond to a size of
the emitting area of 40\,kpc. It might correspond to the X-ray
outflow detected by \citet{Bogdan:2008} (see Figure \ref{fig:XHa},
left panel).} (2) As discussed by various authors
\citep[e.g.][]{David:2006,Ho:2009}, the ionised gas in galaxy bulges
can be accounted for by mass loss of evolved stars and SNIa could play
a key role. The prompt SNIa component (which represents 50\% of the
SNIa population) discussed by \citet{Mannucci:2006} could hence be
indirectly linked to the 100-200Myr star formation episodes quoted
above { and contribute to this outflow}.
As discussed in the following, a possible collision with M32 can
explain the ring structures of M31 where the CO has been detected, and
such an event could have triggered a star formation activity in the
central part and support the SNIa explanation for the outflow. In
addition, this inner outflow could be connected with the outflow
detected in X-ray by \citep{Bogdan:2008} on a larger scale and
discussed in Sect. \ref{ssect:other}.
\begin{figure*}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig8a.eps}
\includegraphics[width=6.5cm,angle=-90]{fig8b.eps}
\caption{Superposition of the various gas velocities (see
Fig. \ref{fig:HaNIIvel} bottom left) (left:) on a simple model of a
galactic disc with an inclination of 45$\deg$, a position angle of
{40$\deg$} and a systemic velocity of -310\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi; (right:)
similar disc with variable position angles} \label{fig:XHamod}
\end{figure*}
Several (8/14) of the positions with line splittings have a component
with a large $[$O{\sc iii}$]$/H$\beta$ ratio. Such a large ionisation is
compatible with planetary nebulae. \citet{DelBurgo:2000} observed
several high-ionisation ``clouds'' in this area. These authors
discussed that the intensity of 3 of these sources is much larger than
those of planetary nebulae detected in M31 by \citet{Ciardullo:2002}.
However, in this central region, it might be possible that these sources
are multiple. Interestingly, one of the sources (D) detected by
\citet{DelBurgo:2000} exhibits a line splitting
(-270,-527)\,km\,s$^{-1}$ with an amplitude comparable with ours but
very close to the centre, even though none of the velocities really
match. Also, the strongest component we detect in the inner arcsec
region matches in first approximation with the source A of
\citet{DelBurgo:2000}.
\begin{figure}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig9.eps}
\caption{Image of the HI emission in the very centre, avoiding
$\pm$100km/s around the systemic velocity of -310km/s. This cut
has been made to subtract the large-scale HI emission seen in
projection superposed to the centre, expected near the systemic
velocity. The contours are the dust emission from {\em
Spitzer}-IRAC { (8$\mu$m image where a scaled version of the 3.6
$\mu$m image has been subtracted to remove the stellar photospheres
emission)} \protect\citep{Block:2006}. The HI cube is from Braun et
al (2009). } \label{fig:HI-dust}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig10.eps}
\caption{Velocity field of the HI emission, corresponding to
{Figure \ref{fig:HI-dust},} avoiding large-scale HI emission.
} \label{fig:velo-both}
\end{figure}
{A comparison with SAURON data (M. Sarzi, private communication)
taken in this region of M31 has shown that our multiple lines
detections are compatible with planetary nebulae. The slit most
probably biases the line ratio determination, as different parts of
the PSF are sampled.}
\subsubsection{Modelling of a tilted inclined disc}
We create a velocity field map with the CCDVEL task within the NEMO
software package \citep{Teuben:1995}. Following
\citet{Ciardullo:1988}, we consider a tilted ring model with an
inclination of 45$\deg$, a position angle of {40$\deg$
\citep{Boulesteix:1987}} and a systemic velocity of -310\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. In
order to understand the behaviour of the gas in the central part of
M31, we superimpose the velocities measured in this paper (ionised gas
and CO gas) on this modelling in Figure \ref{fig:XHamod} {(left
panel)}. Even though there is a velocity field, as shown with ionised
gas by \citet{Boulesteix:1987} and displayed in the bottom right panel
of Figure \ref{fig:XHaboul}, it is not regular and does not exhibit a
well-defined zero-velocity curve. In addition, the velocities of the
CO measurements do not fit with such a regular speed pattern. Varying
the velocity angle in the centre, {as displayed in the right panel
of Figure \ref{fig:XHamod},} mimics the velocity pattern in the
inner region ($<10$\,arcsec) but does not explain the minor axis
configuration.
{\subsection{HI emission} The detailed maps from \citet{Braun:2009}
have revealed clearly the HI deficiency in the central parts of M31.
There is, however, some HI emission in the central kiloparsec, but
most of it comes near the systemic velocity, and is likely to be
projected emission from the external tilted gas orbits, which are
warped to an almost edge-on inclination \citep[e.g.][]{Corbelli:2010}.
To subtract this projected large-radii emission, and better see the
residual coming from the actual centre, with a large velocity
gradient, we have summed all channels avoiding $\pm$100km/s around the
systemic velocity of -310km/s. A weak signal can then be
distinguished, with a morphology of an incomplete ring {(Figure
\ref{fig:HI-dust})}, with emission corresponding to the east side of
the inner ring delineated by the dust emission revealed in the {
8$\mu$m} {\em Spitzer}{-IRAC} map. Figure \ref{fig:HI-dust} shows the dust
contours superposed on the residual high-velocity central HI emission.
In this picture, we can see that the main HI residual component still
follows the large-scale (NE-SW) arms seen in projection, and
coinciding with the dust features, in the direction of the major axis
of M31. However, there remains a weak component perpendicular (NW-SE)
to it. The main concentration of this residual HI emission is well
aligned with the north-east part of the dust ring and it could
correspond to the weak component seen by \citet{Brinks:1983} on the
minor axis position velocity diagram (his figure 1b). However, the
remaining parts of the ring is devoid of HI emission; the atomic gas
must have been transformed into the molecular phase in the dense parts
of the ring. { As discussed in Appendix \ref{sect:append}, Figure
\ref{fig:HIvel}} reveals that most of the CO strong emission at high
velocity in the centre has no HI counterpart.
The HI component associated to the dust ring is, however, too weak to
be seen in the velocity map, as shown from Figure
\ref{fig:velo-both}. The velocity field in the HI selected with high
speed with respect to the systemic velocity, is still compatible to
the ``normal'' {main} disc, and similar to the velocity field
found with the ionised gas in H$\alpha$, although with {possible}
perturbations in the vicinity of the ring. {It is difficult to
determine an accurate position angle of the HI isovelocity map (Figure
\ref{fig:velo-both}) to discriminate between the
nuclear disc and the main disc, but the intensity map (Figure
\ref{fig:HI-dust}) does not really favour a nuclear disc component.}
\subsection{Other wavelengths information}
\label{ssect:other}
\begin{figure*}
\centering
\includegraphics[width=6.5cm,angle=-90]{fig11a.eps}
\includegraphics[width=6.5cm,angle=-90]{fig11b.eps}
\caption{Left: Superposition of the H$\alpha$+$[$N{\sc ii}$]$ map
(upper left panel of figure \ref{fig:HaNIIvel}) of
\citet{Ciardullo:1988} on the Chandra soft X-ray emission from
\citet[see their Figure 7]{Bogdan:2008}. The H$\alpha$+$[$N{\sc
ii}$]$ map has been smoothed with the same smoothing lengths as
those used for the Chandra map (A. Bogdan, private
communication). Right: Superposition of the $A_B$ extinction on the
8$\mu$m {\em Spitzer}{-IRAC} map after subtraction of a scaled version
of the 3.6 $\mu$m image { \citep{Block:2006}}. The $A_B$
contours are fixed to 0.05 (blue) and 0.15 (white).}
\label{fig:XHa}
\end{figure*}
While the centre of M31 hosts a massive black hole with a mass of
$0.7-1.4 \times 10^8 M_\odot$ \citep{Bacon:2001, Bender:2005}, it is
one of the most underluminous supermassive black hole
\citep{Garcia:2010}. The A-star cluster, detected in the { third}
component { (P3) of M\,31's nucleus by \citet{Bender:2005}}, can be
associated to a recent star formation episode, { typically} a
single burst, which occurred 200\,Myr ago, with a total mass in the
range 10$^4$ -- 10$^6 M_\odot$, corresponding to an accretion rate of
10$^{-4}$ -- 10$^{-2} M_\odot$\,yr$^{-1}$. It might be triggered by
the possible frontal collision with M\,32 \citep{Block:2006}. The
detection of an ionised gas outflow in X-rays along the minor axis of
the galaxy by
\citet{Bogdan:2008}, perpendicular to the main disc, could be linked
to this recent star formation activity and to the possible outflow
detected in the optical (Figure \ref{fig:HLS}). However, this is still
controversial as such a burst does not have enough energy to power the
galactic wind required \citep{Bogdan:2008}. In the left panel of
Figure \ref{fig:XHa}, H$\alpha$+$[$N{\sc ii}$]$ contours are
superimposed (with the same resolution) on the Chandra soft X-ray
emission map from \citet{Bogdan:2008}. The relative intensity of the
outflow on both sides is compatible with the intensity of A$_B$
extinction: the NW side is more extinguished than the SE side. This is
compatible with the modelling described in the next section. Last, the
right panel of Figure \ref{fig:XHa} displays superimposition of the
contours of the A$_B$ extinction map on the PAH-dust emission at
8\,$\mu$m detected by {\em Spitzer}. The extinction features match
exactly the ring detected at 8\,$\mu$m.
\section{Interpretation}
\label{sect:inter}
{The molecular, atomic and ionised gas exhibit different radial
distributions in disc galaxies as first discussed by
\citet{Kennicutt:1989} \citep[see also][for a recent
review]{Bigiel:2008}. The HI gas is known to extend at much larger
radius. M31 has HI gas extending at least up to 40\,kpc
\citep{Corbelli:2010,Chemin:2009} with high velocity clouds up to
50\,kpc \citep{Westmeier:2008}. As discussed below, the projected warp dominates
the HI emission in the central part (see also
\ref{sect:append}). The CO and ionised gas are usually more
concentrated. In M31, while CO is known to be depleted in its centre
\citep{Nieten:2006}, the ionised gas is detected in the inner
parts \citep{Ciardullo:1988,Boulesteix:1987} and is strongest there
\citep{Devereux:1994}. }
{While the atomic and ionised gas exhibit the presence a perturbed
disc, the molecular gas detected in CO displays unexpected kinematic
signatures, with significant line splits close to the minor axis and a
very weak when present signal close to the systemic velocity. Several
scenarios could be invoked to explain the existence of two well
separated high S/N velocity components ($\Delta$V = 260\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi
on a scale of 40\,pc) traced by the molecular component (CO emission)
in the centre of M31. One component is in the sense of the expected
rotation, the other component is counter-rotating. The widths of these
two components are also very different, as displayed in Figures
\ref{fig:M31all} and \ref{fig:superimp}. }
The ionised gas also exhibits unexpected kinematic features with a
face-on outflow in the circumnuclear region (detected in the optical),
extending (in Xray) to the 1.5\,kpc $\times$ 1.5\,kpc region studied
in this paper.
In the following, we explore several possible scenarios to
explain the observations. In Sect. \ref{ssect:warploop}, we
discuss the expected effects of a large scale warp. In
Sect. \ref{ssect:bar}, we summarise the expected signatures of a bar
and demonstrate that they fail to match the CO observations. In
Sect. \ref{ssect:ring}, we remind the scenario of a head-on collision
with M\,32 as proposed by \citet{Block:2006} and show with a simple
modelling that it can account for the various observables.
\label{ssect:scenario}
\subsection{Scenario 1: large scale structures warp and tidal streams}
\label{ssect:warploop}
We discuss the arguments that the second velocity component
detected in CO could be due to the superimposition of gas at large
scale onto the central regions. These possible configurations are not
related to the outflows detected there in the ionised gas.
\subsubsection{Large scale warp}
{We could think that the second component, the counter-rotating one,
is just observed in projection, and comes from the external warp
observed in the outer parts of the M31 disc, with a different
inclination and position angle than the normal M31 disc. The warp is
visually conspicuous, and remarkable in HI
\citep[][]{Corbelli:2010}. Farther in the north-east disc (beyond
10\,kpc), \citet{Casoli:1988} already noted that CO emission could
come from the main disc and the warped component, both on the same
line of sight. In that case, the CO emission ($T_A^* \sim 40mK$) was
coming from material distant by more than 16\,kpc from the centre.
This scenario is, however, impossible, because (1) the expected
molecular gas has negligible emission at the large radii where this
warped component should be \citep[e.g.][]{Neininger:1998}; (2) the
corresponding projected velocity on the line of sight close to the
centre is expected around the systemic velocity. It is likely that the
gas at long distance (16-30\,kpc) is on nearly circular orbits and has
no radial velocity when projected to the centre. Accordingly, the
corresponding warped material was always found at systemic
velocity. Whatever inclination it has, the gradient in velocity across
a small region of 40pc should be negligible. Ad hoc hypotheses with HI
infall/outflow with important radial velocity are then required, which
are incompatible with the HI observations (e.g. Corbelli et al
2010). We therefore think that, unless the gas is completely out of
equilibrium
\citep[which is not really supported by the observations of][]{Corbelli:2010,Casoli:1988}, the peculiar component cannot come
from the outer warp. However, we have some tentative detections close
to the systemic velocity (M31K, M31I, M31G, M31J), which could be
accounted for by the external warp. Some positions (e.g. M31G) suffer
from signal subtraction caused by the off positions (wobbler mode of
observation), and we cannot exclude that the real signal could be a
bit larger.}
\subsubsection{Large-scale tidal streams}
\label{ssect:loop}
{The velocity range measured in CO is comparable with the
measurements performed by \citet{Ibata:2004,Chapman:2008} for stellar
streams and for extra-planar gas and high-velocity clouds detected in
HI outside the disc
\citep{Braun:2004,Westmeier:2008}. \citet{Casoli:1988} have detected CO
up to 16\,kpc. \citet{Braun:2004} detected a faint bridge of HI
emission which appears to join the systemic velocities of M\,31 with
that of M\,33 and continues beyond M\,31 to the north-west. Davies's
cloud \citep{Westmeier:2008} lies in the north-west part of M31's disc
and could belong to this bridge. Similarly to the stellar streams
observed in the south-east part of M31's disc \citep{Ibata:2007}, the
HI gas exhibits large design loop-like features \citep[e.g. the
Magellanic stream,][]{Kalberla:2006}. The second component detected in
the molecular gas we have detected could be associated to the M31-M33
bridge or any gaseous-loop relics. It would then correspond to some
accretion of gas onto the disc, generating an inner polar ring and
leading to compression and excitation of the CO. This configuration
can be compared to the inner polar gaseous disc discussed by
\citet{Sil'chenko:2011} in NGC\,7217, where the disc is face-on.
While HI studies mention contamination by the Galaxy
\citep{Westmeier:2008}, our CO detections are related to extinction
patterns that are clearly associated to M\,31. The CO detected for M31D has a
mean velocity of -74\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi is typically below the limit for Galactic
contamination used for HI. However, the typical velocity dispersion of HVC
in the Galaxy is 8.5\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, while the dispersion measured for M31D is
24\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. Furthermore, it exhibits a higher than expected
I$_{CO}$/E(B-V) ratio, rather suggesting that it lies behind the bulge
and not in front of it. Finally, no CO from the Milky Way has been
detected by \citet{Dame:2001} in the vicinity of M31.
Last, we cannot exclude that some non-circular velocities detected in
CO result from relics of M31's formation: scattered debris or other
extra-planar gas.
}
\subsection{Scenario 2: a possible bar}
\label{ssect:bar}
Another a priori plausible scenario could be that the second
velocity component is associated to a possible bar. Bars are frequent
in spiral galaxies \citep[60 to 75\% are barred according to the bar
strength; e.g.][]{Verley:2007}, and it is likely that M31 is
barred { or has been barred, because of its triaxial boxy bulge
\citep{Beaton:2007}. However, up to now, the bar is not visible in the gas
in either the morphological or kinematical data.}
\subsubsection{Previous bar propositions in M31}
\label{ssect:bar1}
There have been several propositions of bar models for M31. However,
they are all contradictory (orientation of the bar along the minor
axis or the major axis), and there is no coherent model that fits the
observations. \citet{Stark:1994} assume the existence of a Ferrers bar
potential in the centre to explain some non-circular motions there,
but there was no complete velocity field, and they do not try to fit
the observed stellar distribution with a bar. Their main point is to
propose that the morphology of the gas in the centre of M31 looks
round, { in spite of the bar, because of the high (77$^{\circ}$)
inclination of M31: the bar would be aligned along the minor axis. The
{\it x1} orbits, elongated along the bar, i.e. the minor axis, are
then projected as almost round, because their axis ratio could be as
high as 4.44= 1/cos(77$^{\circ}$)}. The authors propose {\it x1}
orbits between 0.7 and 3\,kpc in radius, and perpendicular {\it x2}
orbits inside 0.7\,kpc. The gas would be shocked at the transition
between {\it x1} and {\it x2} orbits, in a ring of 0.7\,kpc radius.
However, as is demonstrated below {in Sec. \ref{ssect:bar2}}, this
scenario cannot account for the large observed velocity splitting
along a single line of sight on the minor axis, at $\sim$ 3.5kpc from
the centre.
\citet{Athanassoula:2006} investigate the twist of isophotes of the
NIR 2MASS image, which shows a boxy shape bulge. They compare this
morphology with simulated stellar models with bars, without gas. They
conclude that the NIR morphology of M31 in the centre favours a bar,
seen almost side-on, i.e. the angle between the bar and the major axis
of the galaxy is between 20 and 30$^{\circ}$. The size of the
boxy-bulge is 1.4\,kpc in radius, and the total length of the bar would
be 4\,kpc in radius, because a thin bar is identified beyond the boxy
bulge \citep{Beaton:2007}.
Let us note that in these two above propositions, the bar almost
end-on, or almost side-on, there is no large angle between the bar
and the symmetry axes of the projected galaxy (major or minor
axis). This is required because of the morphological symmetry of the
disc (no bar is obvious in the gas), and also because of the fairly
regular velocity field. If the bar were inclined by $\sim$ 45$^\circ$
with respect to the major axis, strongly skewed {(S-shape)} {
isovelocities} should be seen in the gas, i.e. the kinematical minor
axis would not be perpendicular to the major axis, as in N2683
\citep{Kuzio:2009}, which has been compared to M31. In the latter, owing
to the relative deficiency of gas in the centre
\citep{Brinks:1984b,Nieten:2006}, it is difficult to trace {
isovelocities}, but Figs. \ref{fig:XHaboul} and
\ref{fig:velo-both} reveal that there is no strong skewness.
\citet{Athanassoula:2006} also {present} a 2D gaseous velocity field in
an analytical bar potential (not fitted to M31) {and project it}
edge-on, {i.e. all regions of the galaxy up to 20\,kpc can thus be
seen artificially on the same line of sight, which makes a big
difference from a 77$^\circ$ projection corresponding to M31's real
configuration.} The model 2D gas morphology clearly shows a bar, with
emptied regions that are not similar to the de-projected view of M31
gaseous disc \citep[e.g.][]{Braun:1991}. The authors compare the
position-velocity diagrams parallel to the major axis that they
derived from the 2D-edge-on model to those observed in HI. The latter
reveal conspicuous figure-eight shapes that turn at the end of the HI
extent (R=70\arcmin $\sim$ 15\,kpc), and have been interpreted as
caused by the warp \citep{Henderson:1979,Brinks:1984a}. {These
features} are much larger than the {(3-4\,kpc)} bar extent. Note that
in the galaxy NGC 2683, as inclined on the sky plane as M31, the
position-velocity diagram along the major axis is not a figure-eight
shape, but a parallelogram that ends at a certain radius, precisely
the radius of the bar R= 2.2\,kpc= 45\arcsec
\citep{Kuzio:2009}. The situation is completely different in M31.
\subsubsection{Special case of M31}
\label{ssect:bar2}
Is it possible to obtain two velocity components on each side of the
systemic velocity along the same line of sight, near the minor axis
(position of M31-I and J), at a distance of 210\,arsec, i.e. 3.5\,kpc
from the centre (if the points are in the main inclined plane)? The
possibility to gather several components in the same beam strongly
depends on the spatial resolution. Here, the CO(2-1) beam is 12\arcsec
$\sim$ 40\,pc on the major axis, however, the inclination of the plane
limits the resolution. Typically, gaseous discs in the centre of
galaxies have a characteristic scale-height of 50\,pc, as determined
in the Milky Way \citep[e.g.][]{Sanders:1984}. The inclined line of
sight across the disc of M31 encompasses a region of size
=tan(77$^\circ$) $\times$ 50\,pc= 216\,pc in the direction parallel to
the minor axis (while it is the size of the beam, 40\,pc, in the
perpendicular direction). The region explored in one line of sight
then goes along the minor axis from 3.4\,kpc to 3.6\,kpc, and only
40\,pc in the perpendicular direction. The only place where it would
be possible to have two velocity components on each side of the
systemic velocity is the very centre, because the 216\,pc region
explored { would encompass the two sides of an elongated orbit.}
Assuming that the bar is inclined with some angle with respect to the
major axis, the elongated orbits will provide red- and blue-shifted
velocities on each side of the centre. This is the case in NGC2683 for
instance, which is revealed by a long slit exactly along the major
axis. It is possible in the centre, because the bar is inclined with
respect to the symmetry axes. However, at a distance of 3.5\,kpc from
the centre ($\pm 100\,pc$), it is impossible to obtain these two
V-components, with the same category of orbits, either {\it x1} or
{\it x2}. The only other possibility so far from the centre would be
to consider a change of orientation in the orbits, or a sudden shock,
as occurs from {\it x1} to {\it x2}, or crossing an arm. However, the
amplitude of the shock, 260\,km\,s$^{-1}$, and the positions of the
two components on the two sides of V$_{sys}$ is unrealistic,
especially for a possible weak bar.
\subsubsection{Examples of bar shocks}
\label{ssect:bar3}
Several strongly barred galaxies have been studied in H$\alpha$
spectroscopy, determining the kinematics with high spatial resolution,
comparable to what we have in the CO gas towards M31. NGC\,1365 is
the prototype of a very strong bar. The position angle of the bar does
not coincide with the symmetry axis of the projection, so that the bar
signatures are obvious. With an inclination of i=42$^\circ$, and a
distance of D=20\,Mpc, the observed velocity gradient across the bar
dust lanes is determined to be at maximum 40\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, with a spatial
resolution below 100\,pc \citep{Lindblad:1996}. If deprojected, the
maximum would be 60\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. NGC\,1300 is another remarkable and strongly
barred spiral galaxy, where the kinematics has been studied in detail
in H$\alpha$, on scales smaller than 100\,pc. \citet{Lindblad:1997}
present a projected velocity gradient across the dust lanes with
shocks of 20-30\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi projected on the sky. When deprojected, this
becomes a maximum of 35-50\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. To reach these strong velocity
gradients, a highly contrasted gas structure must be found, which is
not the case for the M31 central region.
In the molecular component, interferometric observations of nearby
barred galaxies can now yield sub-arcsec resolution, corresponding to
scales smaller than 100\,pc, therefore comparable to the single-dish
resolution towards M31. \citet{Garcia-Burillo:2005} have presented
non-circular motions caused by bars, which are of the same orders of
magnitude $\sim$ 60-70\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. More recently, with even higher
sensitivity and resolution, \citet{vanderLaan:2011} have for the first
time seen dedoubled velocity profiles in the CO data towards the
barred galaxy NGC 6951. If interpreted in terms of crowding of orbits,
they reveal velocity differences of 80\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi {(within 500\,pc)},
which in deprojection would amount to 110\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. The CO velocity
profiles are, however, very broad in the centre of this galaxy,
showing a starburst ring, and it is not sure what is caused by turbulence
or by the bar. The molecular gas surface density in the centre is on
average 100\,M$_\odot$/pc$^2$ {(corresponding to a total mass of
$1.5 \times 10^9\ifmmode{{\rm M}_{\odot}}\else{M$_{\odot}$}\fi$ inside 2\,kpc)}, which may explain the unstable and
turbulent gas-forming stars, while in M31 the surface density is at
least one order of magnitude lower. The velocity dispersion in the
distinct velocity components observed in CO is quite low (15\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi),
which corresponds to a stable gas layer. We conclude that the bar is
not a likely explanation for the special kinematical phenomenon
observed towards the M31 centre.
\subsubsection{Other interpretations}
\label{ssect:bar4}
From the HI-derived gas morphology,
\citet{Braun:1991} suggests a two-armed trailing density wave with a
pattern speed of 15\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\,kpc$^{-1}$, leading to a corotation
at 16\,kpc, an ILR at 5\,kpc, and an OLR at 22\,kpc. He proposes as the
driver of the spiral structure the collision with M32, which could
also be responsible for the many tilts of the M31 plane. The nearly
head-on collision with M32 has been developed in detail by
\citet{Block:2006}, which allows them to account for the two
off-centred gas rings observed in M31 at 0.7\,kpc and 10\,kpc
radius. Indeed, these authors note that the gas morphology in M31 does
not show the usual signatures of the response to a bar. Although it is
possible that the galaxy disc possesses a stellar bar, aligned with
the triaxial bulge, oriented roughly along the major axis, the
observed 0.7\,kpc dust ring is aligned almost along the minor axis,
and not aligned with the potential bar along the major axis. The size
of the inner ring does not correspond to a possible inner Lindblad
resonance, nor the outer ring at 10\,kpc radius to a correlated outer
resonance. Moreover, the inner disc appears highly inclined with
respect to the main disc, and the inner ring is off-centred by 40$\%$
of its radius, which is not expected in barred galaxies. This offset
centre and the tilted plane of the inner disc strongly point towards a
perturbed origin of the inner features of M31.
\subsection{Scenario 3: head-on collision with M32}
\label{ssect:ring}
\citet{Jacoby:1985} {have shown} that the inner disc of M31 is
tilted with respect to the main disc. While the inclination of the
large-scale disc is 77$^\circ$, the nuclear disc, of size $\sim$
1\,kpc, is inclined by only about 40$^\circ$ on the sky plane.
\citet{Block:2006} discovered in the {\em Spitzer}-IRAC maps an
inner dust disc of scale 1\,kpc by 1.5\,kpc, which also appears with
this low inclination. The ring morphology of this tilted structure
together with the ring (10kpc) morphology of the large-scale disc led
these authors to propose a head-on collision with a small companion, with a
probable candidate being M32.
In this scenario, we assume that the peculiar component comes from a
tilted ring-like material, likely coming from the perturbed gas from
the recent M32 collision. The collision might have perturbed the
nuclear disc, which was already tilted, and produced some local warp
before the annular density wave propagated outwards. {Besides its
interest for the inner configuration of M31, this scenario also
proposes an explanation for the 10\,kpc ring, resulting from the
propagation of the initial annular wave.}
The resulting configuration is similar to the scenario proposed in
Sect. {\ref{ssect:loop}}. Below, we discuss some modelling that
enables one to explain the observations. We stress that this modelling
is meant to be schematic rather than reproducing the observations in a
fully self-consistent manner.
We used a model for the gas, through static simulations of dynamical
components represented by particles, with different geometrical
orientations on the sky, but embedded in a gravitational potential
representing the observed rotation curve of M31.
We represent the gaseous disc of M31 by a fairly homogeneous
Miyamoto-Nagai disc of particles (with radial scale of 1\,kpc, and
height of 0.2\,kpc) to be able to vary easily the inclination and
position angle as a function of radius. We used typically half a
million particles to have sufficient statistics. We plunged the gas disc
into a potential made of a stars and a dark matter halo. The stellar
component is composed of a bulge and a disc. The bulge is initially
distributed as a Plummer sphere, with a potential
\begin{equation}
\Phi_{b}(r) = - { {G M_{b}}\over {\sqrt{r^2 +r_{b}^2}} },
\end{equation}
where $M_{b}$ and $r_{b}$ are the mass and characteristic radius of the
bulge, respectively (see Table \ref{condini}).
The stellar disc is initially a Kuzmin-Toomre disc of surface density
\begin{equation}
\Sigma(r) = \Sigma_0 ( 1 +r^2/r_d^2 )^{-3/2}
\end{equation}
with a mass $M_d$,
and characteristic radius $r_d$.
The dark matter halo is also a Plummer sphere, with mass $M_{\rm DM}$
and characteristic radius $r_{\rm DM}$. A summary of the adopted
parameters is given in Table \ref{condini}. The resulting rotational
velocity curve is rising until a maximum of 300\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, reached at
3\,kpc, and then slightly decreases, remaining close to 300\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\, in
the central regions of interest here. This corresponds to the
rotational velocity observed
\citep[e.g.][]{Carignan:2006}. All molecular clouds we aim to
reproduce are inside the radius of maximum rotational velocity.
The gas particles are distributed in this potential, with the velocity
dispersion corresponding to a Toomre $Q$ parameter of $1.3$.
\begin{table}[ht]
\caption[ ]{Model parameters}
\begin{flushleft}
\begin{tabular}{ccccccc} \hline
Component & radius & Mass & Mass fraction \\
& [kpc] & [$M_\odot$]& [ \%] \\
\hline
Bulge & 0.2 & 1.6e10 & 4.5 \\
Disc & 4.0 & 7.e10 & 20. \\
Halo & 10. & 27.e10 & 75.5 \\
\hline
\end{tabular}
\end{flushleft}
\label{condini}
\end{table}
\begin{figure}[ht]
\centering
\includegraphics[angle=-90,width=8.5cm]{fig12.eps}
\caption{ Mean density of CO emission from the homogeneous nuclear disc plus tilted ring of
the model. The field of view is 200 arcsec in radius, or 0.76\,kpc in radius. The density scale, indicated in the wedge, is in arbitrary units.
{The positions of our CO observations are indicated with circles corresponding to the beams.}}
\label{mean}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[angle=-90,width=8.5cm]{fig13.eps}
\caption{{ Density-weighted mean velocity (first moment)} of the
simulated nuclear region in the same field of view
as Fig \ref{mean}. Signatures of the tilted ring can be seen at the
boarder of the field. The wedge gives the velocity scale in \ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi.}
\label{velo}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[angle=-90,width=8.5cm]{fig14.eps}
\caption{{ Density-weighted velocity dispersion (second moment of
the simulated velocity field).} The wedge gives the velocity scale in
\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi. The locations of double components with counter-rotating
features correspond to the blue and purple regions. }
\label{width}
\end{figure}
To reproduce the observations, we adopted an inclination of 43$^\circ$
for the gaseous nuclear disc, inside the 1.5\,kpc radius. At the
boundary of this disc, we assumed a progressive warping of the disc
plane, so that the inclination on the sky grows from 43 to
77\,degrees, over 300\,pc. The details of this transition, however,
are not constrained by the observations, because the latter are all
inside the 1\,kpc radius. As for the position angle on the sky, the
nuclear disc has PA$=70^\circ$, unlike the main disc, which has
PA$=35^\circ$. In projection on the sky, the nuclear gas disc model
gives velocities that agree with the observations, at least with the
main velocity component. In many observed points only this main
component is observed, with a broad line-width (FWHM$=$50-70\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi),
located on the blueshifted side in the SW and on the redshifted side
in the NE of the major axis. In some of the points there is an
additional peculiar velocity component, located on the opposite side
(redshifted in the SW), and narrower (20 \ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi).
Trying several parameters, we found that material on a ring-like
orbit, with 40$^\circ$ inclination, and PA= -35$^\circ$ gives
kinematical features compatible with the observations (cf Figure
\ref{mean}-\ref{spec}). The modelled ring has a width of 0.4\,kpc, and
a mean radius of 0.6\,kpc, coinciding with the observed dust ring. The
adopted position angle and inclination of the various components are
displayed in Table \ref{geometry}.
\begin{table}[ht]
\caption[ ]{Geometrical parameters}
\begin{flushleft}
\begin{tabular}{ccccccc} \hline
Component & Main disc & Nuclear disc & Inner ring \\
& [$^\circ$] & [$^\circ$]& [ $^\circ$] \\
\hline
\hspace{0.2cm}
PA & 35 & 70 & -35 \\
Inclination & 77 & 43 & 40 \\
\hline
\end{tabular}
\end{flushleft}
\label{geometry}
\end{table}
\begin{figure}[ht]
\centering
\includegraphics[width=8cm]{fig15.eps}
\caption{Spectra extracted from the simulated cube at the (RA, DEC)
offset in arcseconds indicated in each panel. The vertical scale is
in arbitrary units. The velocities are centred at the systemic
velocity. These offsets correspond the M31G and M31I positions.
The clumpiness of the gas (not included in the modelling proposed
here) is expected to modulate the relative intensity of the
various components additionally.}
\label{spec}
\end{figure}
To draw these maps, we built data cubes corresponding to the
observations, with a pixel of 6 arcsecond, and channels of 12 \ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi,
and the data smoothed to a beam of 12\arcsec. The {pixel} size of the
cubes are therefore (60, 60, 40) to describe the inner parts studied
here in CO lines and dust extinction. To take into account that the
gas and dust distribution is not homogeneous, but patchy, we smoothed
the dust map obtained by \citet{Block:2006} from {\em Spitzer}-IRAC
images (the 8$\mu$m\, map with the 3.6$\mu$m\, contribution subtracted) to
the same spatial resolution, and used it as a multiplicative filter
for our homogeneous disc map. This takes into account that the dust
inner ring is offset by 0.5\,kpc (131'') from the centre of the
galaxy.
Typical spectra were extracted in the zone where double features are
expected at the offset indicated in Fig \ref{spec} in arcseconds. On
the one hand, the broad characteristics of the blue-shifted velocity
component is caused by beam smearing of the velocity gradient, as
illustrated here with a 12$\arcsec$ beam. On the other hand, because
the ring is narrow in radius ($\sim$ 170\,pc) and in a region of a
flat rotation curve, the velocity gradient in the beam is very flat
and a narrow line is observed in the redshifted component.
In addition, some emission at the systemic velocity is expected
because of the inclination of the 77$\deg$ main disc. It is therefore
not necessary to evoke the large-scale warp to explain the weak
emission at the systemic velocity, which can be accounted for by the
inner main disc. For positions like M31I/M31J/M31G located in the
north-west ring, main disc regions at 3.5\,kpc project in the same
beam.
\section{Discussion and conclusions}
\label{sect:disc}
\begin{figure} \centering
\includegraphics[height=7cm]{fig16.eps}
\caption{Schematic view of the interpretation proposed for the CO velocities
observed. The inner disc is presented with a PA of 70$\deg$ and an
inclination of 43$\deg$. The inner ring is superimposed with a similar
inclination but a position angle of -35$\deg$. The straight line
indicates the position of the major axis of the main disc inclined by
77$\deg$ with a PA of 35$\deg$. The bars labelled ''N'' and ''E''
have a length of 150\,pc. { The near (resp. far) sides of the two
components are coloured (resp. kept white). The red (resp. blue)
colours correspond to the redshift (resp. blueshift) relative to the
systemic velocity.}}
\label{fig:schem}
\end{figure}
We have detected CO emission, tracing the presence of molecular gas in
several positions in the central kpc of M31, corresponding to dust
extinction. {We have also shown that there is an excellent
correspondence between the dust extinction measured in the optical and
the dust emission in the near-infrared at 8$\mu$m (right panel of
Figure \ref{fig:XHa}).} Some of the {observed} positions correspond to
the conspicuous inner ring, detected in dust emission by IRAC on {\em
Spitzer} \citep{Block:2006}}. This ring was also present in ISO data
studied by \citet{Willaime:2001} and in ionised gas maps as proposed
by \citep{Jacoby:1985}. The velocity information brought by the CO
lines reveals a complex kinematical structure, different from the
ionised {and atomic gas dynamics}. In the NW inner ring positions, two
{main high S/N} velocity components are detected on each side of the
systemic velocity. The component with the expected velocity,
according to the rotation curve and the positive position angle, is
broad, and is likely to correspond to the nuclear disc. Exploring a
wide region of the strong central rotational gradient, the
beam-averaged spectrum may be 100km/s broad. The second component,
redshifted on the other side of the systemic velocity, has a small
velocity width, because it is not a full disc, but a relatively narrow
ring. We used radii from 0.4 to 0.8\,kpc for the model.
We summarise in Figure \ref{fig:schem} the geometry { of the two
components} we propose in our modelling. The peculiar component
appears to be counter-rotating, because the gas is in a tilted ring,
almost perpendicular to that of the inner disc. Its inclination on the
plane of the sky is similar, which explains why the amplitude of the
projected rotation is the same as for the regular component. In the
figure the arrows indicate the apparent sense of rotation. The two
discs appear to rotate in the same direction, which is also that of
the main M31 disc. The apparent counter-rotation appears only in the
NW (and SE) regions, where blue and red regions superpose. The
winding sense of the spiral structure in the observations is also
sketched. It is possible to see that the arms are trailing, exactly
like the M31 main disc.
In summary, the apparent counter-rotation is only caused by a warping
and distortion of the central components, possibly triggered the
proposed head-on collision with M32, but is not a true
counter-rotation.
\begin{acknowledgements}
We are most grateful to F. Viallefond, who designed the observing
procedures and to M. Marcelin, who retrieved for us old kinematical
$[$N{\sc ii}$]$ data. We thank R. Saglia who has kindly provided the
slit spectra he took in the bulge, enabling us to search for double
components and {\'A} Bogd{\'a}n for providing their reduced XMM-Newton
and Chandra maps of the bulge exhibiting an X-ray outflow and helpful
comments. We acknowledge I. Chilingarian for his help in the spectra
reduction and M. Sarzi for helpful comments. {We are very grateful to
Robert Braun for his constructive comments, and for having provided
the HI data, in order to check the HI correlation with the CO data.}
We thank the anonymous referee for his constructive remarks, which
helped us to improve the manuscript. This paper is based on
observations carried out with the IRAM 30m telescope. IRAM is
supported by INSU/CNRS (France), MPG (Germany) and IGN (Spain). The
authors are grateful to the IRAM staff for their support.
\end{acknowledgements}
|
\section{Introduction}
Studies of central gas and stellar kinematics of nearby galaxies have
revealed the presence of supermassive black holes (SMBHs) in almost all
massive galaxies. According to the most accepted paradigm (Rees 1984),
galactic nuclear activity is directly related with
the fueling process of these SMBHs, which strongly correlates to the overall
structure and evolution of the galaxies. For instance, active galactic nucleus (AGN) activity seems
to be triggered only in galaxies with massive bulges and intermediate
colors, as pointed out by Choi et al. (2009), henceforth CWP09.
These authors
show how the necessary conditions for AGN activity depend on galaxy
morphology; for early-type galaxies low velocity dispersion and
blue colors are preferred, and indicative of available gas supply for
the central black hole, while intermediate velocity dispersion,
intermediate-color, and more concentrated late-type galaxies are
more likely to host AGNs, reinforcing the idea that AGN host galaxies
are intermediate objects between the red sequence and the blue cloud
(Kauffmann et al. 2003).
If most galaxies present SMBHs in their centers,
a fundamental question is why not all of them host AGNs
when gas supply is available.
This is a question with no simple answer given the
complexity of the processes involved in the formation and evolution
of galaxies, but important clues have been proposed to elucidate the
mechanisms that trigger AGN activity (Martini 2004). Major mergers of gas-rich
progenitors and galaxy interactions are two commonly invoked mechanisms
that can provide gas to the center of the galaxies and fuel dormant
SMBHs (Hopkins et al. 2006; Ellison et al. 2008). Tidal torques
exerted by non-axisymmetric features such as large-scale and nuclear
bars (Jogee 2006), recycled gas from dying stars in the inner
kiloparsecs of the galaxy (Ciotti \& Ostriker 2007), and enhanced viscosity
of the turbulence associated with supernova explosions (Chen et al. 2009) are just
some other mechanisms that can effectively trigger nuclear activity.
A common requirement for any of the mentioned mechanisms is the removal of
specific angular momentum in order for the material to collapse to the
center, and operate as fuel. Therefore, a relation between the overall
galactic $\lambda$ spin parameter and the onset of AGN activity is expected,
with a higher frequency of AGNs in low spinning galaxies. Statistically,
AGN hosting galaxies should form a low $\lambda$ population, in comparison
to normal galaxies. Low angular momentum is expected not only in the
growing phase of SMBHs, but has also been invoked to explain the formation
of massive black hole seeds themselves. In this scenario
(Eisenstein \& Loeb 1995; Koushiappas et al. 2004), the direct
collapse of low angular momentum material results in a typical inverse
proportionality between the SMBH mass and the total angular momentum
(Haehnelt et al. 1998; Colgate et al. 2003).
In this work, we study the spin distribution
of AGN host galaxies in contrast with non-AGNs, parameterizing the
angular momentum through the dimensionless spin parameter
\begin{equation}
\label{Lamdef}
\lambda = \frac{L \mid E \mid^{1/2}}{G M^{5/2}},
\end{equation}
where $E$, $M$ and $L$ are the total energy, mass and angular momentum of the configuration,
respectively (Peebles 1969). Our aim is to determine empirically if the AGN population
effectively presents a low spin distribution when compared with non-AGNs
and corroborate the results arising
from theoretical studies concerning the relation between $\lambda$ and the SMBH mass.
To this end, we use a sample of galaxies selected from the Sloan
Digital Sky Survey (SDSS) and apply a simple model to obtain
an empirical estimation of $\lambda$. The large volume limited sample
allows a thorough coverage of parameter space, which permits an empirical
determination of the leading galactic physical parameters driving the AGN
phenomena and the observed scalings with SMBH mass. The physical model is
presented in \S~2, and the sample selection criteria in \S~3. In
\S~4 we present our general results and the final conclusions
are summarized in \S~5.
\section{Estimation of the spin from observable parameters}
In Hernandez \& Cervantes-Sodi (2006), we derived a simple estimate of $\lambda$
for disk galaxies in terms of observational parameters
and showed some clear correlations between this parameter
and type defining structural parameters.
Here we briefly recall the main ingredients of the simple
model. The model considers only two components, a disk for
the baryonic component with an exponential surface mass density $\Sigma(r)$,
\begin{equation}
\label{Expprof}
\Sigma(r)=\Sigma_{0} e^{-r/R_{\rm d}},
\end{equation}
where $r$ is a radial coordinate, and $\Sigma_{\rm 0}$
and $R_{\rm d}$ are two constants which are allowed to
vary from galaxy to galaxy, and a dark matter halo having
a singular isothermal density profile $\rho(r)$,
responsible for establishing a rigorously flat rotation
curve $V_{\rm d}$ throughout the entire galaxy:
\begin{equation}
\label{RhoHalo}
\rho(r)={{1}\over{4 \pi G}} \left( {{V_{\rm d}}\over{r}} \right)^{2}.
\end{equation}
In this model we are further assuming that
(1) the specific angular momentum of the disk and halo are equal
(e.g. Mo et al. 1998; Zavala et al. 2008); (2) the total energy is dominated
by that of the halo which is a virialized gravitational structure; and
(3) the disk mass is a constant fraction of the halo mass,
$F=M_{\rm d}/M_{\rm H}$. These assumptions allow us to express $\lambda$ as
\begin{equation}
\label{Lamhalo}
\lambda=\frac{2^{1/2} V_{\rm d}^{2} R_{\rm d}}{G M_{H}}.
\end{equation}
Finally, we introduce a baryonic Tully--Fisher (TF) relation:
$M_{\rm d}=A_{\rm TF} V_{\rm d}^{3.5}$ (e.g., Gurovich et al. 2004).
Taking the Milky
Way as a representative example, we evaluate $F$ and $A_{\rm TF}$ to obtain
\begin{equation}
\label{LamObs}
\lambda=21.8 \frac{R_{\rm d}/\rm kpc}{(V_{\rm d}/\rm km \: s^{-1})^{3/2}}.
\end{equation}
We proved the accuracy in our estimation of $\lambda$ in Cervantes-Sodi et al. (2008),
comparing the estimation using
Equation (\ref{LamObs}) to values arising from numerical simulations of six distinct groups,
where the actual value of $\lambda$ is known, as it is one of the parameters
of the simulated galaxies, and where this can also be estimated through Equation (\ref{LamObs}),
as baryonic disk scale lengths and disk rotation velocities are part of the output.
The result was a one-to-one correlation, with very small dispersion and no bias,
leading to typical errors $<$ 30\%, which we include throughout.
The use of this estimate for large SDSS samples
led us in Hernandez et al. (2007) and Cervantes-Sodi et al. (2008) to a first empirical
derivation of the distribution of $\lambda$ parameters in the local universe
and a measure of the scalings of this parameter with local environment and halo mass.
These results were latter confirmed by Berta et al. (2008) using an estimate of
$\lambda$ through Equation (\ref{LamObs}), also used by Gogarten et al. (2010)
as an unbiased first-order $\lambda$ measurement.
\section{The SDSS sample}
The sample of galaxies used in this work is an enhanced version
of the study by CWP09.
The galaxies come from the large-scale structure
sample, DR4plus, using the New York University
Value-Added Galaxy Catalogue (VAGC; Blanton et al. 2005), which is a galaxy subset
of the SDSS Data Release 5 (Adelman-McCarthy et al. 2006), and the
Korea Institute for Advanced Study-VAGC (Choi et al. 2010). A detailed
description of the sample can be found in Choi et al. (2007) and CWP09.
The rest-frame absolute magnitudes of individual galaxies were computed in
the $r$ band, using Galaxy reddening correction (Schlegel et al. 1998) and
$K$-corrections as described by Blanton et al. (2003), with a mean evolution
correction given by Tegmark et al. (2004).
We define a volume-limited sample with a lower magnitude cut-off
$M_{r} = -19.0$, with redshifts $0.020 < z < 0.086$, and a total number of
87,590 galaxies. Disk scale lengths are estimated from one-component light fits,
which implicitly account for the presence of low angular momentum material
in bulges.
\begin{figure}
\centering
\begin{tabular}{c}
\includegraphics[width=0.475\textwidth]{fig1.eps}
\end{tabular}
\caption[ ]{Galaxy distribution in the BPT diagram, pure AGN galaxies lying above the solid
line (Kewley et al. 2006), star-forming galaxies below the dashed line (Kauffmann et al. 2003),
and composite objects in between.}
\label{Fig1}
\end{figure}
In order to compare the spin of AGN and non-AGN galaxies, we use only Type II
AGNs to avoid complications due to the presence of a bright nucleus. The AGN
galaxies were segregated from star forming galaxies using the flux ratio
BPT diagram of Balmer and ionization lines (Baldwin et al. 1981).
The AGNs were selected among galaxies fulfilling the definition by
Kewley et al. (2006):
\begin{figure*}
\label{distributions}
\centering
\begin{tabular}{cc}
\includegraphics[width=0.475\textwidth]{fig2.eps} & \includegraphics[width=0.475\textwidth]{fig4.eps} \\
\includegraphics[width=0.475\textwidth]{fig3.eps} & \includegraphics[width=0.475\textwidth]{fig5.eps}
\end{tabular}
\caption[ ]{(a) $P(\lambda)$ distribution for late-type
galaxies in the sample segregated in normal,
composite and AGN galaxies. (b) $P(\lambda)$ distribution
for normal galaxies split according to absolute magnitude.
(c) $P(\lambda)$ distribution
for AGN galaxies split according to absolute magnitude.
(d) Median $\lambda$ value of AGN galaxies and the control sample
as a function of $M_{\rm r}$.}
\end{figure*}
\begin{eqnarray}
0.61/({\rm log}([\rm N_{II}]/H\alpha)-0.47)+1.19&<&{\rm log}([\rm O_{III}]/H\beta),\nonumber\\
0.72/({\rm log}([\rm S_{II}]/H\alpha)-0.32)+1.30&<&{\rm log}([\rm O_{III}]/H\beta),\nonumber\\
0.73/({\rm log}([\rm O_{I}]/H\alpha)+0.59)+1.33&<&{\rm log}([\rm O_{III}]/H\beta)\nonumber,
\label{eq2}
\end{eqnarray}
while star-forming galaxies were selected using the selection criteria by Kauffmann
et al. (2003):
\begin{eqnarray}
0.61/({\rm log}([\rm N_{II}]/H\alpha)-0.05)+1.30&<&{\rm log}([\rm O_{III}]/H\beta),\nonumber\\
0.72/({\rm log}([\rm S_{II}]/H\alpha)-0.32)+1.30&<&{\rm log}([\rm O_{III}]/H\beta),\nonumber\\
0.73/({\rm log}([\rm O_{I}]/H\alpha)+0.59)+1.33&<&{\rm log}([\rm O_{III}]/H\beta), \nonumber
\end{eqnarray}
requiring a signal-to-noise ratio $\geq$ 6.
The galaxies residing in the area between the previous demarcation
lines (solid and dotted lines, respectively, in Figure~\ref{Fig1})
are identified as composite objects, which contain AGN as well as extended H$_{\rm II}$
regions. Our non-AGN sample consist of those galaxies excluding AGNs and composite
galaxies, without signal-to-noise cut.
To estimate the black hole mass ($M_{BH}$), we use the $M_{BH}$--stellar velocity dispersion
relation by Tremaine et al. (2002):
\begin{equation}
\label{BHmass}
{\rm log} (M_{\rm BH}/M_{\odot}) = 8.13 + 4.02 \times {\rm log} (\sigma/200~{\rm km~s}^{-1}),
\end{equation}
where a simple aperture correction to the stellar velocity dispersion was applied. Since velocity
dispersion measurements lower than the instrumental resolution are not robust, we will focus on AGNs
with $M_{\rm BH} > 10^{6.7}M_{\odot}$.
In order to discriminate between elliptical and disk galaxies, we used the prescription of
Park \& Choi (2005) in which early (ellipticals and lenticulars) and late (spirals) types are
segregated in a $u - r$ color versus $g - i$ color gradient space and in the concentration index space.
They extensively tested the selection criteria through a direct comparison of visually assigned types
for a large sample of several thousand galaxies. The specific selection criteria can be found in Park \& Choi (2005).
For more details the reader is referred to CPW09.
\section{Results}
Our Equation (\ref{LamObs}) requires the rotational velocity to calculate $\lambda$,
which is inferred from the absolute
magnitude in the $r$-band and using a TF relation (Pizagno et al. 2007).
To avoid the problem of internal
absorption in edge-on galaxies (Cho \& Park 2009), and consequently underestimating rotational
velocities, we limit the sample to spiral galaxies having seeing-corrected
isophotal axis ratios $b/a > 0.6$.
Once the galaxies in our sample are segregated into early- and late-type galaxies,
and into different activity classes such as normal, composite, and AGN hosting
galaxies, we computed $\lambda$ using Equation (\ref{LamObs}) for all
the late-type galaxies and obtained the spin distribution for the three
different subclasses: 41,662 normal, 2,273 composite and 1,026 AGNs.
Figure 2(a) shows the distribution
of $\lambda$ values for normal, composite, and AGN late-type host
galaxies. We can see a clear difference between the normal and the
AGN population, the former showing typically higher spin values with a
large dispersion compared with the latter, which is a more coherent
population of low spin galaxies.
The distribution of galaxies classified as composite follows almost exactly
the same distribution as AGN galaxies, probably due to the low number of
misclassified star-forming galaxies. Given that these galaxies are actually
a mixed group of AGN and non-AGN galaxies we will not consider them for the
rest of the analysis, and we will focus only in the difference between AGNs and
normal galaxies.
Theoretical (Shaw et al. 2006), as well as empirical (Hernandez et al. 2007)
distributions of $\lambda$, are commonly described by a log-normal
function, in our case, the values describing the non-AGN and AGN distributions are, respectively,
$\lambda_{0}=0.059\pm0.006$, $\sigma_{\lambda}=0.400\pm0.005$ and $\lambda_{0}=0.042\pm0.004$, $\sigma_{\lambda}=0.350\pm0.003$,
showing a clear difference between the two populations.
We use the two sample Komogorov--Smirnov test to compare the distributions of normal and
AGN host galaxies to determine if they are drawn from the same parent distribution.
The result of the test
yields a negligible probability of $P \sim \times 10^{-180}$
that the two samples are drawn form the same parent distribution.
If we look at the $\lambda$ distribution of normal galaxies, splitting the sample
according to the absolute magnitude (Figure 2(b)),
we clearly see that the distributions get tighter and present
lower mean $\lambda$ values
for brighter galaxies, the same behavior we reported
in Cervantes-Sodi et al. (2008), where we found that massive galaxies tend to
have low spin dark matter halos and show low dispersion, while low
mass galaxies have higher $\lambda$ values and high dispersions around the median
value, a result later confirmed by Berta et al. (2008).
A striking difference appears for the case of AGN host galaxies
(Figure 2(c)), showing low mean $\lambda$ values
and tight distributions for all $M_{r}$. This result indicates
that only low spinning galaxies, at all mass ranges,
are suitable AGN hosts in the local universe.
Previous work has already pointed out that the recurrence of AGN activity is a
function of internal properties such as absolute magnitude, color, and velocity
dispersion. Kauffmann et al. (2003) state that massive galaxies with old stellar
populations frequently host AGNs and
CWP09 found that the fraction of AGN increases with galaxy luminosity,
and particularly for late-type galaxies,
this fraction increases for intermediate color systems,
with $u$ -- $r = 2\sim2.4$ and intermediate velocity dispersion ($\sim130$ km s$^{-1}$).
In this context, most of our AGN hosting galaxies are restricted to
having intermediate colors, a restriction that also constrains the
value of $\lambda$ given that color anticorrelates with $\lambda$ (Hernandez
\& Cervantes-Sodi 2006).
To make a fair comparison between AGN and non-AGN galaxies,
we constructed a control sample by selecting non-AGN galaxies
which match one to one the morphology, absolute magnitude, color and concentration
index of each galaxy in the AGN sample, guaranteeing that the
distributions of these parameters are the same for both populations.
Figure 2(d) shows how the typical $\lambda$ value of AGN hosting
galaxies tends to be lower than the one presented by the control sample
at all $M_{r}$ values. The uncertainties (throughout this paper) represent 1$\sigma$ confidence
intervals determined by the bootstrap resampling method.
\begin{figure}
\begin{tabular}{c}
\label{fractions}
\includegraphics[width=.475\textwidth]{fig6.eps} \\
\includegraphics[width=.475\textwidth]{fig7.eps}\\
\includegraphics[width=.475\textwidth]{fig8.eps}
\end{tabular}
\caption[ ]{AGN fraction as a function of(a) absolute magnitude with cuts of $\lambda$
and as a function of $\lambda$
with cuts in magnitude (b) and color (c).}
\end{figure}
A natural question arising from our previous results is
whether or not the AGN fraction depends on galactic spin.
Previous studies have shown the dependence of the AGN fraction
on the brightness of the parent galaxies (Mercurio et al. 2010, CWP09), or their associated
stellar or total mass (Pasquali et al. 2009). Figure 3(a)
shows the fraction of AGN+composite ($f_{\rm AGN+Composite}$)
galaxies as a function of absolute magnitude in different $\lambda$ cuts,
where a moderate increase of $f_{\rm AGN+Composite}$ is present for increasing brightness,
while a large difference in this fraction is evident for the distinct $\lambda$ cuts.
Figure 3(b) gives a complementary study
for $f_{\rm AGN+Composite}$ as a function of galactic spin for thin $M_{r}$ slices,
where the dependence on $\lambda$ appears stronger than the case of $M_{r}$,
with all curves consistent with a single underlying dependence to within errors,
for all magnitude bins. This shows that the driving physical ingredient is
the value of $\lambda$ rather than magnitude.
CPW09 have already shown that late-type galaxies present weak dependence of $f_{\rm AGN+Composite}$
on magnitude or mass (accounted for through velocity dispersion) in addition to $u$ -- $r$ color, and
in the preceding section we found that the value of $\lambda$ is intimately linked to
the color of the galaxy in question. The $f_{\rm AGN+Composite}$ dependency on $\lambda$ for narrow
$u$ - $r$ bins is presented in Figure 3(c), where even at fixed color, a substantial
decrease of the AGN fraction results as $\lambda$ increases, at least in the
range of colors where the highest occurrence of AGNs is found. We see that regardless of mass or color, high
$\lambda$ systems will only very rarely present an AGN.
The tight correlations of $M_{\rm BH}$ with host galaxy properties, such as the
well-known one with the bulge luminosity (Kormendy \& Richstone 1995) or the tighter
$M_{\rm BH}$ versus bulge velocity dispersion relation (Graham et al. 2011),
have now been extended to more global
properties, not related to the bulge component alone.
Correlations with the spiral arm pitch angle of disk galaxies (Seigar et al.
2008), the mass of the dark matter halo (Ferrerase 2002; Booth \& Schaye
2010), and the total gravitational mass of the host galaxies (Bandara
et al. 2009) show how the correlations are not restricted only
to the bulge component and point to a SMBH--bulge--total-mass interrelation (Volonteri et al. 2011).
Given the important role the spin plays in establishing the morphology
and present-day structure of disk galaxies, and our previous result
that shows the direct influence of $\lambda$ on determining nuclear
activity, it is logical to expect a correlation between $\lambda$ and
$M_{\rm BH}$ for the active galaxies in our sample;
more specifically, an anticorrelation.
Let us consider the simple criterion by Larson (2010) for the formation of black holes,
where gas in a forming galaxy with a column density above a critical value $\Sigma_{\rm crit}$,
goes into forming a central black hole. If the gas in the forming galaxy is distributed
radially like the total mass with an isothermal density profile given by Equation
(\ref{RhoHalo}), with a gas to total mass ratio $f_{\rm g}$, the surface
density of gas at any radius $r$ in the disk will be $\Sigma_{\rm g}(r) =
f_{\rm g} V_{\rm d}^{2}/ 2 \pi G r$. For a rotationally supported disk, the scale
radius is given by $R_{\rm d} \simeq \lambda r_{\rm vir}$,
where $r_{\rm vir}$ is the virial radius of the
system. This allow us to express the mass of gas within the critical column density criterion as
\begin{equation}
M_{g}(\Sigma < \Sigma_{\rm crit} )= \frac{f_{g} M^{2}}{\pi \Sigma_{\rm crit} (\lambda r_{\rm vir})^{2}}.
\label{MBH-lambda}
\end{equation}
This expression, that does not attempt to be more than a dimensional analysis estimate,
implies that at fixed total mass the mass of gas that could end up forming a black
hole is $M_{\rm BH} \propto \lambda^{-n}$, with $n=2$.
Theoretical studies report different values
for $n$ that fall in the range 1--4 (Colgate et al. 2003; Koushiappas et al.
2004), all of them establishing an anticorrelation between the mass of the
black hole and halo spin. Figure 4 (top panel) shows the empirical
dependence of the inferred $M_{\rm BH}$ with the spin parameter for our sample of AGN host
galaxies, showing the expected anticorrelation with $\lambda$, in this case with a best
fit described by log($M_{\rm BH}$) = ($-0.716 \pm 0.114$) log($\lambda$) $+ (6.455 \pm 0.155$).
In Figure 4 (bottom
panel) we present the result of splitting the sample according to the absolute magnitude of the
galaxies, where we can appreciate how at fixed $\lambda$ brighter (more massive) galaxies host
more massive black holes, with the anticorrelation with $\lambda$
present at all fixed $M_{r}$ cuts. In this case, the leading effect is the total
mass, with $\lambda$ introducing only a small but well-defined trend in the expected sense.
\begin{figure}
\centering
\begin{tabular}{c}
\label{l-mbh}
\includegraphics[width=0.475\textwidth]{fig9.eps}
\end{tabular}
\caption[ ]{Upper panel:
$M_{\rm BH}$ versus $\lambda$ for AGNs hosted by late-type galaxies, showing
the median value for each bin. Lower panel: the same as in the upper panel
but showing the median $M_{\rm BH}$ values splitting the sample according to
the absolute magnitude.}
\end{figure}
\section{Conclusions}
Using an extensive sample of galaxies extracted from the SDSS DR5, we show that
the empirical distributions of the $\lambda$ spin parameter of AGNs and
non-AGN host galaxies are qualitatively and quantitatively different, the first distribution
having lower $\lambda$ values and smaller dispersion around the median
when compared with the latter. Another striking difference between the two populations is
that while the spin of normal galaxies presents a dependence on absolute magnitude, with
a systematic decrease of $\lambda$ with increasing luminosity, AGN host galaxies
present at all $M_{r}$ low $\lambda$ values. This result, in addition to the increase
of the AGN fraction for decreasing $\lambda$, highlights the requirement
for galaxies to have low spin in order to host AGNs and should be taken into account
in semi-analytic models that exclusively use the halo or stellar mass to reproduce
the fraction of galaxies belonging to different activity classes (e.g., Wang \& Kauffmann 2008;
Fontanot et al. 2011). Seeding AGNs into modeled dark matter halos should primarily consider
$\lambda$, in order to better model the observed universe.
For the AGN sample we found that the inferred mass of the SMBHs shows a weak dependence on
the spin of the hosting galaxies, with increasing mass for decreasing spin, a logical
result if we need the gas to be accreted onto the central massive object,
a process that can be more effectively accomplished if the raw material originally
has low angular momentum. We recovered the leading $M_{\rm BH}$-$\lambda$ relation at fixed
absolute magnitude, with typically higher SMBH mass for brighter galaxies at fixed
$\lambda$, showing a double dependence of $M_{\rm BH}$ on $M_{r}$ and $\lambda$.
Given that the trend with the spin is present at all $M_{r}$ bins, this could
account, to a certain degree, for the dispersion in the well established
$M_{\rm BH}$--luminosity relation.
We thank the anonymous referee for a thorough reading and constructive comments that helped to
enrich our study. B.C-S. thanks Cheng Li for useful comments.
Y.Y.C. was supported by grant from the Kyung Hee University (KHU-20100179)
and the National Research Foundation of Korea to the Center for Galaxy Evolution Research.
|
\section{Introduction}
The two-particle-irreducible (2PI) effective action \cite{LW,LY,Baym1,Baym2,Cornwall:vz} offers a general and systematically improvable approach for resumming infinite classes of Feynman diagrams of a given quantum field theory. One of its most compelling aspects is that it appears to be applicable to a wide variety of situations ranging from out-of-equilibrium settings to finite temperature calculations in equilibrium. In fact, the first orders of approximation of the 2PI effective action seem to capture already many interesting features of quantum field theories. In the case of scalar fields for instance, including the first non-local, field independent contribution to the 2PI effective action is enough to obtain a controlled (non-secular) time evolution which shows thermalization at late times, at least at the level of the two-point function \cite{Berges:2000ur}. In equilibrium, the same orders of approximation lead to a rather good convergence of some thermodynamic quantities, such as the pressure \cite{Berges:2004hn}, indicating that part of the infrared physics is properly taken into account. These observations are not limited to scalar theories but concern also theories coupling scalar and fermionic degrees of freedom (out-of-equilibrium) \cite{Berges:2002wr} or gauge theories (at finite temperature) \cite{Blaizot:2000fc,Blaizot:2005wr}. Some more genuine non-perturbative effects can be captured by combining the 2PI effective action with a large-N type expansion \cite{Berges:2001fi,Aarts:2002dj}. In the non-equilibrium context the 2PI 1/N expansion at next-to-leading order proved to be particularly fruitful for the study of a variety of problems, such that thermalization \cite{Berges:2001fi,Aarts:2002dj,Cooper:2002qd}, preheating \cite{Berges:2002cz,Arrizabalaga:2004iw}, transport coefficients \cite{Aarts:2003bk,Aarts:2004sd,Aarts:2005vc}, non-thermal fixed points \cite{Berges:2008wm}, decoherence \cite{Giraud:2009tn} and topological defect formation \cite{Berges:2010nk}. In equilibrium, the 2PI formalism was applied to phenomenological studies in various approximations, see \cite{Petropoulos:2004bt} and references therein as well as \cite{Andersen:2008qk,Roder:2005vt}. Of course, one can easily imagine that certain non-perturbative features are beyond reach within the 2PI framework or its present approximations. However, higher order approximations of the 2PI effective action or its generalization to 3PI, 4PI, ..., NPI effective actions \cite{DeDom}, although numerically unaffordable at present, could allow to capture an increasing number of such effects in the future \cite{Carrington:2004sn,Berges:2004pu,Carrington:2010qq}. Some investigations in these directions can be found in \cite{Aarts:2006cv,Carrington:2009zz,Carrington:2009kh}.\\
Concomitantly with the increasing number of applications of the 2PI effective action, new insight has been gained on technical aspects regarding its renormalization. The systematic renormalization of approximations based on the 2PI effective action was first understood in the case of scalar theories at finite temperature both in the real-time formalism \cite{vanHees:2001ik,VanHees:2001pf} and in the imaginary time formalism \cite{Blaizot:2003an,Berges:2005hc}. It was then extended to the case of scalar theories coupled to fermions \cite{Reinosa:2005pj} and also to abelian gauge theories \cite{Reinosa:2006cm}. Another important extension consisted in including the possibility of non-vanishing field expectation values, of relevance for the determination of the effective potential and thus the study of phase transitions \cite{vanHees:2002bv,Berges:2005hc,Cooper:2005vw,Arrizabalaga:2006hj,Patkos:2008ik,Fejos:2009dm}. The renormalization of models with more complicated global symmetry was studied in \cite{Fejos:2007ec}. From all these studies, a certain consensus has emerged that the renormalization of approximations based on the 2PI effective action seems always possible provided certain extensions of the renormalization procedure are allowed in order to cope with certain approximation artifacts. For instance, in the presence of non-vanishing field expectation values, it is well known that there exist different expressions for the two- and four-point functions, and more generally for higher $n$-point functions. Although equivalent in the exact theory, these various expressions differ within a given approximation and bring their own divergences which can only be absorbed by allowing for apparently more counterterms than usual \cite{Berges:2005hc}. Some of these counterterms are fixed by means of the usual renormalization conditions. The others are fixed by imposing consistency conditions, that is conditions which would be satisfied automatically if no approximation was considered at all \cite{Berges:2005hc,Reinosa:2009tc}. The fact that the consistency conditions do not involve the parameters of the theory is crucial to maintain the number of such parameters to its expected value, despite the larger number of counterterms. In fact, these apparently different counterterms should be viewed as different subseries of the complete perturbative counterterm series and thus should agree when no truncation of the 2PI effective action is considered.\\
In this paper we illustrate partly these issues regarding renormalization by revisiting one of the simplest approximations of the 2PI effective action at non-zero field and finite temperature in the case of a real scalar field, the so-called Hartree approximation. The calculation of the finite temperature effective potential of a real scalar $\varphi^4$ theory in the Hartree approximation and the order of the phase transition that it predicts have been discussed many times in the literature \cite{Espinosa:1992gq,AmelinoCamelia:1992nc,Verschelde:2000ta,Smet:2001un,Cooper:2002qd}. Similar discussions exist for different approximations in the $O(N)$ model with $N\geq 2$, see \cite{Baym:1977qb}. There is a wide consensus on the fact that the order of the phase transition predicted by this approximation is first order, even though all existing studies are either based on numerical evaluations for particular values of the parameters or involve a high temperature expansion which is not justified for all values of parameters. Part of the originality of this work is that, without relying on a high temperature expansion, it presents a complete analytic confirmation of these results, in the whole parameter space.\\
In section~\ref{sec:2PI}, we briefly recall how the effective potential is computed using the 2PI effective action. The main ingredient is the resolution of a self-consistent equation for the two-point function, the so-called gap equation, which encodes the resummation of particular classes of Feynman diagrams. The Hartree approximation corresponds to the leading order of a systematic approximation scheme for evaluating the 2PI effective action and the corresponding gap equation is then an equation for a self-consistent mass. This equation encodes the resummation of superdaisy diagrams \cite{Dolan:1973qd,AmelinoCamelia:1992nc,Drummond:1997cw} and is discussed at length in section~\ref{sec:gap}. We first recall how renormalization of the self-consistent mass is performed at finite temperature and non-vanishing field expectation value, following a slightly different presentation than the one which is usually found in the literature (although equivalent in practice). We put the emphasis on the fact that not all the divergences of the gap equation lead to divergences of the self-consistent mass. The quadratic divergence of the gap equation is indeed responsible for a divergence of the self-consistent mass, which needs to be absorbed in the renormalization of the bare mass, as usual (sections~\ref{sec:mass} and \ref{sec:triviality}). In contrast, the remaining logarithmic divergence of the gap equation does not lead to a divergence of the self-consistent mass. Instead, it makes this self-consistent mass trivial in the continuum limit, that is equal to the renormalized mass, independently of the value of the temperature or the field expectation value. In this respect, the renormalization of the bare coupling does not appear as a way to absorb any divergence in the self-consistent mass but as a way to avoid triviality (sections~\ref{sec:triviality} and \ref{sec:Landau}). As it is well known, the price to pay for defining a non-trivial scalar theory by means of coupling renormalization is that the theory needs to be regarded as a cut-off theory, with the cut-off $\Lambda$ taken below a certain scale, known as the Landau scale $\Lambda_{\rm p}$. It is then important to discuss how calculations can be made insensitive to $\Lambda$ in this context. This is discussed in section~\ref{ss:UVity} where we provide a complete analytical discussion of the solutions of the gap equation for any value of $\Lambda$, below and above the Landau scale. Finally, in section~\ref{sec:flow}, we present a new look at the gap equation based on ``evolution'' or ``flow'' equations for the thermal and field dependence of the self-consistent mass: these equations not only provide an efficient way to solve the gap equation but they also shed a new light on its renormalization. Section \ref{sec:eff_pot} is then devoted to the analysis of the effective potential in the Hartree approximation. Its renormalization is described in sections~\ref{sec:geom} and \ref{sec:ren_eff_pot} with an emphasis on the links with the more general approach presented in \cite{Berges:2005hc} and the distinction between renormalization and consistency conditions. Sections~\ref{sec:extrema} and \ref{sec:temp} discuss analytically the shape of the effective potential as the temperature is lowered from an initial ``high'' temperature $T_\star$ where the theory is chosen to be in the symmetric phase, down to $T=0$. We prove in particular that, depending on the values of the parameters, there is either no transition or a first order phase transition. Section~\ref{sec:concl} is devoted to a discussion concerning the renormalization in the broken symmetry phase and to conclusions.
\vfill
\pagebreak
\section{The 2PI-resummed effective potential}\label{sec:2PI}
In what follows, we consider a real scalar $\varphi^4$ theory in four dimensions at finite temperature, defined by the Euclidean action:
\beq
S[\varphi]\equiv\int_0^{1/T} d\tau\int d^3x\left(\frac{1}{2}(\partial_\tau\varphi)^2+\frac{1}{2}(\nabla\varphi)^2+\frac{m_0^2}{2}\,\varphi^2+\frac{\lambda_0}{4!}\,\varphi^4\right),
\eeq
where the inverse temperature sets the size of the compact interval for temporal integration, and $m_0$ and $\lambda_0$ denote respectively the bare mass and the bare coupling. Notice that, in order that the spectrum be bounded from below, one should restrict to $\lambda_0\geq 0$. In what follows we restrict our attention to the case $\lambda_0>0$.\\
The two-particle-irreducible (2PI) formalism provides a representation of the effective potential $\gamma(\phi)$ corresponding to $S[\varphi]$ in terms of 2PI diagrams. More precisely, it is obtained as the value taken by the 2PI functional:
\beq\label{eq:gen}
\gamma[\phi,G]=\frac{m_0^2}{2}\,\phi^2+\frac{\lambda_0}{4!}\,\phi^4+\frac{1}{2}\int_Q^T\ln G^{-1}+\frac{1}{2}\int_Q^T(Q^2+m_0^2)\,G+\Phi[\phi,G;\lambda_0]
\eeq
at its stationary point $G=\bar G_\phi$, that is $\gamma(\phi)=\gamma[\phi,\bar G_\phi]$ with
\beq\label{eq:stat}
0=\left.\frac{\delta\gamma}{\delta G}\right|_{\phi,\,\bar G_\phi}\,.
\eeq
In Eq.~(\ref{eq:gen}), $\phi$ represents a homogeneous field configuration and $G(i\omega_n,q)$ a function of the Matsubara frequency $\omega_n\equiv2\pi nT$ and the three dimensional momentum $q$. We have also adopted the notation:
\beq\label{eq:notation}
\int_Q^T f(Q)\equiv T\sum_n\int_q\,f(i\omega_n,q)\equiv T\sum_n\int \frac{d^3q}{(2\pi)^3}\,f(i\omega_n,q)\,.
\eeq
Finally, the functional $-\Phi[\phi,G;\lambda_0]$ corresponds to all $0$-leg 2PI diagrams that one can draw in the shifted theory $S[\phi+\varphi]-S[\phi]-(\delta S/\delta\phi)\varphi$ at finite temperature with propagator $G$. This functional cannot be computed exactly. So-called $\Phi$-derivable approximations consist in retaining in $\Phi[\phi,G;\lambda_0]$ only certain classes of diagrams. In this paper we consider the well known Hartree approximation which corresponds to the truncation:
\beq\label{eq:Hartree}
\Phi[\phi,G;\lambda_0]=\frac{\lambda_0}{4}\,\phi^2\int^T_Q G+\frac{\lambda_0}{8}\left(\int_Q^T G\right)^2.
\eeq
According to the above discussion, in order to compute the corresponding effective potential, we first need to determine the stationary propagator $\bar G_\phi$.
\section{Gap equation}\label{sec:gap}
The stationarity condition (\ref{eq:stat}) can be expressed equivalently as
\beq
\bar G^{-1}_{\phi,\,T}(Q)=Q^2+m_0^2+\left.\frac{2\delta\Phi}{\delta G(Q)}\right|_{\bar G_{\phi,\,T}}\,,
\eeq
where $Q^2\equiv\omega_n^2+q^2$. We have used the subscripts $\phi$
and $T$ to stress the fact that the solution $\bar G_{\phi,\,T}$
depends on both the field $\phi$ and the temperature $T$. In what
follows, we shall omit this notation, unless specifically needed. In
the case of the Hartree approximation (\ref{eq:Hartree}), the
propagator takes a very simple form, namely $\bar
G(Q)=1/(Q^2+\bar M^2),$ with the mass $\bar M$ obeying the so-called
gap equation:
\beq\label{eq:gap}
\bar M^2=m_0^2+\frac{\lambda_0}{2}\left[\phi^2+\int_Q^T \bar G\right].
\eeq
We shall explain how to solve this equation later, in Section~\ref{sec:flow}. Before doing so, a fundamental difficulty needs to be bypassed, namely the fact that the gap equation only makes sense within a given ultraviolet regularization. It follows that its solution(s) can depend strongly on the chosen regularization, unless something is done to remove this sensitivity. This is what renormalization is all about in this context.\\
Let us choose for instance a three dimensional, rotation invariant regularization. After performing the Matsubara sum, the sum-integral entering the gap equation reads explicitly:\footnote{Depending on the context, we use the notations $Q<\Lambda$ and $q<\Lambda$ to designate the same three dimensional regularization.}
\beq\label{eq:mat}
\int_{Q<\Lambda}^T \bar G\equiv\int_{q<\Lambda}\frac{1+2n_{\varepsilon_q}}{2\varepsilon_q}\,,
\eeq
where $\varepsilon_q\equiv (q^2+\bar M^2)^{1/2}$ and $n(\varepsilon)\equiv 1/(e^{\beta\varepsilon}-1)$. The term corresponding to $``1"$ in the numerator of Eq.~(\ref{eq:mat}) is particularly sensitive to $\Lambda$. An explicit calculation leads to
\beq
\int_{q<\Lambda}\frac{1}{2\varepsilon_q}=\frac{1}{8\pi^2}\left[\Lambda\sqrt{\Lambda^2+\bar M^2}-\bar M^2{\rm Arcsinh}\left(\frac{\Lambda}{\bar M}\right)\right],\label{eq:vac}
\eeq
which shows that the gap equation (\ref{eq:gap}) possesses terms which depend quadratically and logarithmically on the scale $\Lambda$. What is meant by this is that if we would take $\Lambda$ to infinity, for fixed $m^2_0$, $\lambda_0$, and $\bar M^2$, the right-hand-side of the gap equation would diverge quadratically and logarithmically. However, one should keep in mind that not all these divergences lead to a divergence of the solution of the gap equation.\footnote{It is true that these divergent terms lead to quadratic and logarithmic divergences in the coefficients of the formal perturbative expansion of the solution $\bar M^2$ in powers of $\lambda_0$. However it is not true that they all lead to a divergence of $\bar M^2$, see the subsequent discussion.} The purpose of the remainder of this section is to clarify the connection between the sensitivity of the gap equation to the scale $\Lambda$ and the sensitivity of its solution(s) to this very same scale, and to explain how the later can be removed, or at least substantially reduced.
\subsection{Quadratic divergence}\label{sec:mass}
Let us first prove that the quadratic divergence of the gap equation (\ref{eq:gap}) leads to a divergence of its solution $\bar M^2$. We first prove that there is a unique solution for large enough $\Lambda$ and then that this solution diverges with increasing $\Lambda$. To this purpose, it is convenient to write the gap equation as $0=f_\Lambda(\bar M^2)$ with
\beq\label{eq:deff}
f_\Lambda(M^2)\equiv-M^2+m_0^2+\frac{\lambda_0}{2}\left[\phi^2+\int_{Q<\Lambda}^T G\right]
\eeq
and $G\equiv 1/(Q^2+M^2)$, and study the positive\footnote{We shall only be concerned with positive ($\bar M^2\geq 0$) solutions of the gap equation. For a discussion of negative solutions of the ``explicit'' form of the gap equation at zero temperature, that is the gap equation where the integrals have been performed explicitly for $M^2>0$ and then extended to any value of $M^2$, see \cite{Nunes:1993bk}.} zeros of $f_\Lambda(M^2)$. The derivative of this function with respect to $M^2$ reads
\beq\label{eq:fprime}
f'_\Lambda(M^2)=-1-\frac{\lambda_0}{2}\int_{Q<\Lambda}^T G^2\,,
\eeq
which is always strictly negative because $\lambda_0>0$. It follows that $f_\Lambda(M^2)$ decreases strictly from
\beq
f_\Lambda(0)=m_0^2+\frac{\lambda_0}{2}\left[\phi^2+\int_{Q<\Lambda}^T\frac{1}{Q^2}\right]
\eeq
to $f_\Lambda(\infty)=-\infty$ (the regularized tadpole integral in
Eq.~(\ref{eq:deff}) is suppressed for $M^2\gg\Lambda^2,\,T^2$). Then,
the existence of a solution of the gap equation depends on the sign of
$f_\Lambda(0)$. Even though the parameter $m_0^2$ could be negative,
the quadratic and positive divergence in $f_\Lambda(0)$, given
explicitly in Eq.~(\ref{eq:vac}), implies that there exists a
value of $\Lambda$ above which $f_\Lambda(0)\geq 0$ and the gap
equation admits a unique solution $\bar M^2$. For the same reason,
given a mass $\mu$, there exists a value of $\Lambda$ above which
$f_\Lambda(\mu^2)\geq 0$ and thus $\bar M^2\geq \mu^2$. This shows
that the solution $\bar M^2$ diverges as $\Lambda\rightarrow\infty$,
for fixed $m^2_0$ and $\lambda_0$, as announced.
\subsection{Renormalization of the mass -- Triviality}\label{sec:triviality}
Since the quadratic divergence of the gap equation depends neither on $\phi$ nor on $T$, it can be absorbed by adjusting the divergent part of $m^2_0$. On the other hand, the finite part of $m^2_0$ can be used to impose a condition at a given value of $\phi$ and a given value of $T$. We choose $\phi=0$, $T=T_\star$ and impose the renormalization condition:
\beq\label{eq:rencondm}
\bar M^2_{\phi=0,\,T_\star}=m^2_\star\,.
\eeq
The parameter $m^2_\star$ is positive by construction, because it is a solution of the gap equation at $\phi=0$ and $T=T_\star.$ We shall choose it strictly positive in what follows. The renormalization condition (\ref{eq:rencondm}) is quite natural when studying how the system evolves as the temperature $T$ is decreased from a ``high'' temperature $T_\star$ where the system is required to be in the symmetric phase, see Section~\ref{sec:eff_pot}. It can be rewritten as a choice of the bare mass, namely:
\beq\label{eq:m0}
m_0^2=m_\star^2-\frac{\lambda_0}{2}\int_{Q<\Lambda}^{T_\star} G_\star\,,
\eeq
with $G_\star\equiv 1/(Q^2+m^2_\star)$. With this choice, the gap equation reads $0=\tilde f_\Lambda(\bar M^2)$, with
\beq\label{eq:gapnm}
\tilde f_\Lambda(M^2)\equiv-M^2+m_\star^2+\frac{\lambda_0}{2}\left[\phi^2+\int_{Q<\Lambda}^T G-\int_{Q<\Lambda}^{T_\star}G_\star\right].
\eeq
The dependence of the gap equation on $\Lambda$ has been changed by the renormalization procedure and we
need to reconsider its possible solutions and their dependence on $\Lambda$. As before, we do so by discussing the zeros of the function $\tilde f_\Lambda(M^2)$.\\
Notice first that $\tilde f'_\Lambda(M^2)=f'_\Lambda(M^2)$ is strictly negative in view of Eq.~(\ref{eq:fprime}). It follows that the function $\tilde f_\Lambda(M^2)$ decreases strictly from
\beq
\tilde f_\Lambda(0)=m_\star^2+\frac{\lambda_0}{2}\left[\phi^2+\int_{Q<\Lambda}^T\frac{1}{Q^2}-\int_{Q<\Lambda}^{T_\star}G_\star\right]
\eeq
to $\tilde f_\Lambda(\infty)=-\infty$ (the regularized tadpole
integral in Eq.~(\ref{eq:gapnm}) is suppressed for
$M^2\gg\Lambda^2,\,T^2$). Using Eqs.~(\ref{eq:mat}) and
(\ref{eq:vac}), it is easily checked that $\tilde f_\Lambda(0)$
diverges logarithmically like $c\,m_\star^2\ln\Lambda,$ with
$c>0$. Then, there exists a value of $\Lambda$ above which $\tilde
f_\Lambda(0)\geq 0$ and the gap equation admits a unique solution
$\bar M^2$. For the same reason, for any $\Delta m^2>0$, $\tilde
f_\Lambda(m_\star^2\pm\Delta m^2)$ diverges logarithmically like
$\mp\,c\,\Delta m^2\ln\Lambda$. Then, there exists a value of
$\Lambda$ above which $\tilde f_\Lambda(m^2_\star-\Delta m^2)\geq 0$
and $\tilde f_\Lambda(m^2_\star+\Delta m^2)\leq 0$, and thus $|\bar
M^2-m^2_\star|\leq\Delta m^2$. This proves that $\bar M^2\rightarrow
m^2_\star$ as $\Lambda\rightarrow\infty$, for fixed $\lambda_0$. Using
the gap equation $0=\tilde f_\Lambda(\bar M^2)$ as well as
Eqs.~(\ref{eq:mat}) and (\ref{eq:vac}), it is possible to determine
precisely how this limit is approached. We obtain
\beq\label{eq:approach}
\bar M^2-m^2_\star\sim\frac{8\pi^2}{\ln\Lambda}\left[\phi^2+\int_{q}\frac{\delta_\star n_{\varepsilon^\star_q}}{\varepsilon^\star_q}\right]\,,
\eeq
where $\varepsilon^\star_q\equiv (q^2+m^2_\star)^{1/2}$ and $\delta_\star n_\varepsilon\equiv n_\varepsilon-n^\star_\varepsilon$, with $n^\star_\varepsilon$ the thermal factor at temperature $T_\star$. Thus, unlike what happened with the quadratic divergence, the logarithmic divergence of the gap equation (the one that remains after mass renormalization) does not lead to a divergence of the solution $\bar M^2$.\footnote{The absence of divergence in the solution of the mass renormalized gap equation means that the resummation of superdaisy diagrams encoded in this equation formally resums the perturbative logarithmic divergences associated with each diagram of this series into a convergent contribution. That perturbative divergences can be resummed to non-divergent expressions was observed in \cite{Ghinculov:1997tn,Fejos:2009dm}.} Rather, if we insist in taking the limit $\Lambda\rightarrow\infty$ for fixed $\lambda_0$, the solution of the gap equation converges to the renormalized mass $m^2_\star$ for any value of the field $\phi$ and the temperature $T$. This illustrates the triviality of $\varphi^4$ theory \cite{Drummond:1997cw}, at least in the particular approximation considered here. Equation (\ref{eq:approach}) shows that the trivial limit is approached rather slowly (logarithmically).
\subsection{Renormalization of the coupling -- Landau pole}\label{sec:Landau}
The previous analysis has shown that, for the particular choice of mass renormalization we have considered, the triviality of $\varphi^4$ theory is related to the presence of a logarithmically divergent term in the gap equation. In order to define a non-trivial theory, one can absorb this divergence in a redefinition of $\lambda_0$. Using the formula (see Appendix~\ref{app:id}):
\beq\label{eq:id}
\int_{Q<\Lambda}^T \bar G=\int_{Q<\Lambda}^{T_\star}\bar G+\int_{q<\Lambda}\frac{\delta_\star n_{\varepsilon_q}}{\varepsilon_q}\,,
\eeq
as well as the identity:
\beq\label{eq:id3}
\bar G-G_\star=-(\bar M^2-m^2_\star)\,G_\star\,\bar G=-(\bar M^2-m^2_\star)\,G_\star^2+(\bar M^2-m^2_\star)^2\,G_\star^2\,\bar G\,,
\eeq
it is clear that the logarithmic divergence in Eq.~(\ref{eq:gapnm}) is entirely accounted for by the term $-(\bar M^2-m^2_\star)\int_{Q<\Lambda}^{T_\star}G_\star^2$. Isolating this contribution and defining an effective coupling $\lambda_\star$ \cite{Coleman:1974jh} such that
\beq\label{eq:effl}
\frac{1}{\lambda_\star}\equiv\frac{1}{\lambda_0}+\frac{1}{2}\int_{Q<\Lambda}^{T_\star} G_\star^2\,,
\eeq
the gap equation $0=\tilde f_\Lambda(\bar M^2)$ can be rewritten as $0=g_\Lambda(\bar M^2)$ with
\beq\label{eq:gapr}
g_\Lambda(M^2) & \equiv & -M^2+m^2_\star+\frac{\lambda_\star}{2}\left[\phi^2+\int_{Q<\Lambda}^T G-\int_{Q<\Lambda}^{T_\star}G_\star+(M^2-m^2_\star)\int_{Q<\Lambda}^{T_\star}G_\star^2\right]\nonumber\\
& = & -M^2+m^2_\star+\frac{\lambda_\star}{2}\left[\phi^2+\int_{q<\Lambda}\frac{\delta_\star n_{\varepsilon_q}}{\varepsilon_q}+(M^2-m^2_\star)^2\int^{T_\star}_{Q<\Lambda} G_\star^2\,G\right],
\eeq
where, in the second line, we have made use of Eqs.~(\ref{eq:id}) and (\ref{eq:id3}) to obtain an explicitly convergent expression that we will use later. If we insist in keeping the bare coupling $\lambda_0$ fixed and positive, then, according to Eq.~(\ref{eq:effl}), the effective coupling $\lambda_\star$ goes to zero as $\Lambda\rightarrow\infty$ and we recover the trivial result with $\bar M^2-m^2_\star\sim\lambda_\star\,[\phi^2+\int_q \delta_\star n_{\varepsilon^\star_q}/\varepsilon^\star_q]/2$, in agreement with Eq.~(\ref{eq:approach}). In contrast, fixing the value of $\lambda_\star$ allows to avoid triviality. This corresponds to the following redefinition of $\lambda_0$ ($\lambda_0$ is then $\phi$- and $T$-independent as it should):
\beq\label{eq:l0}
\frac{1}{\lambda_0}=\frac{1}{\lambda_\star}-\frac{1}{2}\int_{Q<\Lambda}^{T_\star} G_\star^2\,.
\eeq
However, it appears that in order to maintain $\lambda_0>0$, and in turn a meaningful microscopic theory, one needs to restrict the cut-off $\Lambda$ below a certain scale known as the Landau scale or Landau pole $\Lambda_{\rm p}$, defined by
\beq\label{eq:Landau}
0=\frac{1}{\lambda_\star}-\frac{1}{2}\int_{Q<\Lambda_{\rm p}}^{T_\star}G_\star^2\,.
\eeq
In other words, the non-trivial Hartree approximation has a meaning only if
it is considered as describing an effective theory. Notice that, from
Eq.~(\ref{eq:Landau}), it follows that $\lambda_\star>0$ (negative values of the renormalized coupling could be possible with other renormalization schemes without violating the requirement $\lambda_0>0$). More explicitly, using Eq.~(\ref{eq:id2}), we obtain
\beq
\frac{1}{\lambda_\star}=\frac{1}{16\pi^2}\left[{\rm Arcsinh}\left(\frac{\Lambda_{\rm p}}{m_\star}\right)-\frac{\Lambda_{\rm p}}{\sqrt{\Lambda^2_{\rm p}+m^2_\star}}\right]+\frac{1}{2}\int_{q<\Lambda_{\rm p}}\frac{n^\star_{\varepsilon^\star_q}-\varepsilon^\star_q{n^\star}'_{\varepsilon^\star_q}}{2{\varepsilon^\star_q}^3}\,.
\eeq
In the limit $\lambda_\star\rightarrow 0^+$, $\Lambda_{\rm p}\rightarrow\infty$ with $\Lambda^2_{\rm p}\sim\mu^2_\star\,e^{32\pi^2/\lambda_\star}$ where $\mu_\star$ depends on $m_\star$ and $T_\star$:
\beq
\mu^2_\star=\frac{m^2_\star}{4}\exp\left(2-16\pi^2\int_q\frac{n^\star_{\varepsilon^\star_q}-\varepsilon^\star_q{n^\star}'_{\varepsilon^\star_q}}{2{\varepsilon^\star_q}^3}\right).
\eeq
\subsection{Ultraviolet sensitivity \label{ss:UVity}}
Because, in the Hartree approximation at least, a non-trivial $\varphi^4$ theory is only to be considered as a
cut-off theory, we need to wonder how calculations done within such a
theory can be made almost independent of the cut-off $\Lambda$. The
main idea is that, if one considers the regime $\Lambda_{\rm p}\gg
T_\star,\,m_\star,\,\phi,\,T$, one can choose the cut-off $\Lambda$
such that both requirements $\Lambda_{\rm p}>\Lambda$ and $\Lambda\gg
T_\star,\,m_\star,\,\phi,\,T$ are met. Then, because all divergences of the
gap equation have been absorbed in the redefinition of the bare
parameters, we expect that its solution $\bar M^2$ be almost cut-off
independent, up to terms of order
$T_\star/\Lambda\,,m_\star/\Lambda\,,\phi/\Lambda\,,T/\Lambda\ll
1$. The Hartree approximation offers the possibility to illustrate this issue, for one can study its solutions as a function of $\Lambda$, in particular as $\Lambda\rightarrow\infty$, and thus assess under which conditions these solutions can be considered practically insensitive to $\Lambda$ for $\Lambda<\Lambda_{\rm p}$.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics[keepaspectratio,width=0.7\textwidth,angle=0]{./sols}
\caption{Number of solutions of the gap equation $0=g_\Lambda(\bar M^2)$
in the $(\Lambda,\phi^2)$-plane in the regime of interest,
that is when $\Lambda,\Lambda_{\rm p}\gg T_\star,m_\star,\phi,T.$
Orange areas correspond to no solution, green areas
correspond to one solution, and the blue area corresponds to two
solutions, as indicated also by the labels. In the regime of interest $\phi^2_{\rm p}\sim\Lambda_{\rm p}^2/(8\pi^2)$ and $\phi^2_{\rm p}(T,\infty)\sim\Lambda_{\rm p}^2/(4\pi^2 e^2)$ are both high scales.\label{fig:sols}}
\end{center}
\end{figure}
The analysis of
the solutions of the gap equation $0=g_\Lambda(\bar M^2)$ as one
varies the cut-off $\Lambda$ is given in Appendix~\ref{app:sol} where, in order to
simplify the discussion, we restrict our analysis to the ``regime of
interest'' that is a regime where both $\Lambda$ and $\Lambda_{\rm p}$
are much larger than all the other scales $T_\star$, $m_\star$, $\phi$,
and $T$. The result of this analysis is that the number of solutions of the gap equation depends on the values of $\Lambda$ and $\phi^2,$ as we represent in Fig.~\ref{fig:sols} and explain in what follows:\\
1. For $\Lambda<\Lambda_{\rm p}$, there is a ``critical'' value
$\phi^2_{\rm c}(T,\Lambda)$ of $\phi^2$ such that the gap equation
admits a unique solution if $\phi^2\geq\phi^2_{\rm c}(T,\Lambda)$ and
no solution if $\phi^2<\phi^2_{\rm c}(T,\Lambda)$, see
Fig.~\ref{fig:sols}. Notice that $\phi^2_{\rm c}(T,\Lambda)$ is not
necessarily positive. In practice it is useful to know when it could
become strictly positive, signaling the fact that the gap equation has no (positive) solutions if $0\leq\phi^2<\phi^2_{\rm c}(T,\Lambda)$. To this purpose, one writes, see Appendix~\ref{app:sol}:
\beq
\phi^2_{\rm c}(T,\Lambda)=-\frac{2\,C_\star}{\lambda_\star}-\int_{q<\Lambda}\frac{n_q}{q}\,,
\eeq
with
\beq\label{eq:Cstar}
C_\star\equiv m^2_\star+\frac{\lambda_\star}{2}\left[-\int_{q<\Lambda}\frac{n^\star_q}{q}+m^4_\star \int^{T_\star}_{Q<\Lambda} \frac{G_\star^2}{Q^2}\right].
\eeq
Clearly, $\phi^2_{\rm c}(T,\Lambda)$ is strictly positive if the parameters $\Lambda$, $T_\star$, $m_\star$, and $\lambda_\star$ are such that $C_\star<0$ and the temperature $T$ is strictly below a certain ``critical'' temperature $T_{\rm c}$ defined by
\beq\label{eq:Tc}
\int_{q<\Lambda}\frac{n^{\rm c}_q}{q}\equiv-\frac{2\,C_\star}{\lambda_\star}\,.
\eeq
Using Eq.~(\ref{eq:Cstar}), it is easy to check that $\phi^2_{\rm c}(T_\star,\Lambda)$ is strictly negative. It follows that $T_{\rm c}<T_\star$. Moreover, in the regime of interest, one can neglect the dependence of $C_\star$ and of the thermal integrals with respect to $\Lambda$. The critical value $\phi^2_{\rm c}(T,\Lambda)$ is then independent of $\Lambda$ (this is the reason why it is represented by an horizontal line in Fig.~\ref{fig:sols}) and one has $\phi^2_{\rm c}(T)=(T^2_{\rm c}-T^2)/12$ with $T^2_{\rm c}=-24C_\star/\lambda_\star$.\\
2. For $\Lambda=\Lambda_{\rm p}$, a new ``relevant'' scale appears:
\beq
\phi^2_{\rm p}\equiv\int_{Q<\Lambda_{\rm p}}^{T_\star}G_\star>0\,,
\eeq
see Appendix~\ref{app:sol}. It is such that $\phi^2_{\rm p}>\phi^2_{\rm c}(T)$ and the gap equation admits no solution if $\phi^2\geq\phi^2_{\rm p}$, one solution if $\phi^2_{\rm p}>\phi^2\geq\phi^2_{\rm c}(T)$, and again no solution if $\phi^2<\phi^2_{\rm c}(T)$, see Fig.~\ref{fig:sols}. Notice that, in the regime of interest, $\phi^2_{\rm p}\sim\Lambda^2_{\rm p}/(8\pi^2)$ is a high scale.\\
3. Similarly, for $\Lambda>\Lambda_{\rm p}$ a new relevant scale $\phi^2_{\rm p}(T,\Lambda)>\phi^2_{\rm c}(T)$ enters the discussion. The gap equation admits no solution if $\phi^2>\phi^2_{\rm p}(T,\Lambda)$ one solution if $\phi^2=\phi^2_{\rm p}(T,\Lambda)$, two solutions if $\phi^2_{\rm p}(T,\Lambda)>\phi^2\geq\phi^2_{\rm c}(T)$ and one solution if $\phi^2<\phi^2_{\rm c}(T)$, see Fig.~\ref{fig:sols}. The value of $\phi^2_{\rm p}(T,\Lambda)$ as $\Lambda\rightarrow\Lambda_{\rm p}^+$ is nothing but $\phi^2_{\rm p}$. Moreover, in the regime of interest, one can show that $\phi^2_{\rm p}(T,\Lambda)$ decreases with $\Lambda$, from $\phi^2_{\rm p}(T,\Lambda^+_{\rm p})=\phi^2_{\rm p}$ to $\phi^2_{\rm p}(T,\infty)\sim\Lambda^2_{\rm p}/(4\pi^2e^2)$, another high scale.\\
The previous analysis shows in particular that the unique solution of the gap equation which exists for $\Lambda<\Lambda_{\rm p}$ and $\phi^2\geq\phi^2_{\rm c}(T)$ and which we shall call the {\it physical solution}, behaves rather differently, depending on the value of $\phi^2$, as one increases the value of $\Lambda$. If $\phi^2\geq\phi^2_{\rm p}$, as we show in Appendix \ref{app:sol}, the physical solution diverges as $\Lambda\rightarrow\Lambda_{\rm p}^-$. In contrast, if $\phi^2_{\rm p}>\phi^2\geq \phi^2_{\rm c}(T)$, nothing of this kind occurs: the physical solution can be extended beyond the Landau scale and a second solution, which we call the {\it unphysical solution}, appears. If $\phi^2_{\rm p}>\phi^2>\phi^2_{\rm p}(\infty)$, there is value of $\Lambda$ at which the two solutions merge into one solution and then cease to exist (without diverging). On the contrary, in the {\it convergence window} $\phi^2_{\rm p}(\infty)\geq\phi^2\geq\phi^2_{\rm c}(T)$ the two solutions can be extended to $\Lambda\rightarrow\infty$ and one can then assess how rapidly they converge to their limiting values $\bar M^2_\infty$. To do so, we take a derivative of the gap equation $0=g_\Lambda(\bar M^2)$ with respect to $\Lambda$. We obtain
\beq
\frac{\partial\bar M^2}{\partial\Lambda}=-\frac{1}{g'_\Lambda(\bar M^2)}\frac{\partial g_\Lambda}{\partial\Lambda}\,,
\eeq
where $\partial g_\Lambda/\partial\Lambda$ corresponds to the explicit dependence of $g_\Lambda$ with respect to $\Lambda$. An implicit dependence is also present, through $\bar M^2$, which explains the appearance of $g'_\Lambda(\bar M^2)$ in the previous formula. We are allowed to divide by $g'_\Lambda(\bar M^2)$ because, as long as $\phi^2<\phi^2_{\rm p}(\infty)$, one has $g'_\Lambda(\bar M^2)\neq 0$, see Appendix~\ref{app:sol}. Moreover, this remain true in the limit $\Lambda\rightarrow\infty$. It follows that
\beq
\frac{\partial\bar M^2}{\partial\Lambda}\sim-\frac{1}{g'_\infty(\bar M^2_\infty)}\left.\frac{\partial g_\Lambda}{\partial\Lambda}\right|_{\bar M^2=\bar M^2_\infty}\,.
\eeq
Then, a simple calculation using Eqs.~(\ref{eq:id1}), (\ref{eq:id2}) and (\ref{eq:un}) leads to
\beq
\frac{\partial\bar M^2}{\partial\Lambda}\sim-\frac{3}{64\pi^2 g'_\infty(\bar M^2_\infty)}\frac{(\bar M^2_\infty-m_\star^2)^2}{\Lambda^3}\,,
\eeq
or, in other words:
\beq\label{eq:convergence}
\left|\frac{\bar M^2}{m_\star^2}-\frac{\bar M^2_\infty}{m_\star^2}\right|\sim \frac{3\lambda_\star}{128\pi^2|g'_\infty(\bar M^2_\infty)|}\left(1-\frac{\bar M^2_\infty}{m_\star^2}\right)^2\frac{m_\star^2}{\Lambda^2}\,,
\eeq
where we have expressed all the masses in units of the renormalized
mass $m_\star$. Although we did not make it explicit, the limiting
value $\bar M^2_\infty$ depends on $T$ and $\phi$. In the case of the physical solution $\bar M^2_\infty$ increases
strictly from $0$ at $\phi^2=\phi^2_{\rm c}(T)$ to a very large value
$\bar M^2_{\rm p}(\infty)\sim 4\Lambda^2_{\rm p}/e^2$ at
$\phi^2=\phi^2_{\rm p}(\infty)\sim\Lambda^2_{\rm p}/(4\pi^2e^2)$, see
Appendix~\ref{app:sol}. Thus, in the regime of interest, such as in
particular $\phi\ll\Lambda_{\rm p}$, $\bar M^2_\infty/m^2_\star$ is of
the order $1$ or smaller and the scale multiplying the convergence
factor $1/\Lambda^2$ in Eq.~(\ref{eq:convergence}) is a low scale of
the order of $m^2_\star\ll\Lambda_{\rm p}^2$ or smaller (the factor
$1/g'_\infty(\bar M^2_\infty)$ can only improve the convegence, for
it goes to zero when $\bar M^2_\infty\rightarrow 0$). Then, the
insensitivity of the physical solution to the scale $\Lambda$ is expected to be observed
already for values of $\Lambda$ below the Landau scale
$\Lambda_{\rm p}$. This expectation is confirmed by the numerical results presented in Fig.~\ref{fig:conv} which shows the convergence of the physical solution as $\Lambda\rightarrow\infty$ for increasing values of $\phi$ : for small values of the field the physical solution has already converged for values
of $\Lambda$ well below the Landau pole, while for very large values
of $\phi$ the convergence, if it happens, occurs for values of $\Lambda$ above the Landau pole. This is a clear illustration of why, although there exists a
Landau pole $\Lambda_{\rm p}$ which forces us to choose the cut-off $\Lambda<\Lambda_{\rm p}$, one can
define, in some region of the parameter space, a theory whose
results are almost insensitive to the cut-off $\Lambda$.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics[keepaspectratio,width=0.495\textwidth,angle=0]{./M_conv}
\includegraphics[keepaspectratio,width=0.495\textwidth,angle=0]{./conv}
\caption{ Convergence of the physical solution of the renormalized and
$\Lambda$-dependent gap equation at different values of the field for
$m_\star^2/T_\star^2=0.5,$ $\lambda_\star=20$ and $T/T_\star=0.012$. The dahed vertical line, common to both plots, indicates the location of the Landau scale. The left panel shows the dependence of the rescaled physical solution with respect to the cut-off $\Lambda$. The right panel illustrates the validity of the asymptotic formula (\ref{eq:convergence}) represented by the solid lines. The inset shows the behavior of the
solution for two values of the field which lie outside of the
convergence window (see text for details).
\label{fig:conv}}
\end{center}
\end{figure}
\subsection{Flow approach to the gap equation}\label{sec:flow}
To conclude this section, we present an elegant way to solve the gap equation. This approach is also interesting on formal grounds because it gives a new perspective on renormalization. It could also be used for higher order approximations.\\
Let us assume that the solution of the gap equation is known for some $\phi$ and $T$, and let us ask how this solution changes as we vary $\phi$ or $T$. The change in the solution is governed by ``evolution'' or ``flow'' equations that we now derive. Let us first express the infinitesimal change of $\bar M^2$ with respect to $\phi^2$. From the original bare gap equation (\ref{eq:gap}) and using the fact that the bare parameters $m^2_0$ and $\lambda_0$ are $\phi$-independent, we obtain
\beq
\frac{\partial \bar M^2}{\partial \phi^2}=\frac{\lambda_0}{2}\left[1+\int_{Q<\Lambda}^T\frac{\partial\bar G}{\partial\phi^2}\right]=\frac{\lambda_0}{2}-\frac{\lambda_0}{2}\int_{Q<\Lambda}^T\bar G^2\,\frac{\partial \bar M^2}{\partial\phi^2}\,,
\eeq
where we used $\delta\bar G=-\bar G^2\,\delta\bar G^{-1}$. Collecting
the derivatives of $\bar M^2$ on the same side of the equation, this can be rewritten as
\beq\label{eq:flowMphi}
\frac{\partial \bar M^2}{\partial \phi^2}=\frac{\bar V}{2}\,,
\eeq
with the function $\bar V$ defined by
\beq\label{eq:BS}
\frac{1}{\bar V}=\frac{1}{\lambda_0}+\frac{1}{2}\int_{Q<\Lambda}^T\bar G^2\,.
\eeq
Similar manipulations are possible to express the infinitesimal change of $\bar M^2$ with respect to $T$. Using the fact that the bare parameters are $T$-independent, one writes
\beq
\frac{\partial \bar M^2}{\partial T}=\frac{\lambda_0}{2}\left(\frac{\partial}{\partial T}\int_{Q<\Lambda}^T\right)\bar G-\frac{\lambda_0}{2}\int_{Q<\Lambda}^T\bar G^2\,\frac{\partial \bar M^2}{\partial T}\,,
\eeq
where the notation $(\partial/\partial T\int_{Q<\Lambda}^T) \bar G$
defined in Appendix~\ref{app:id} refers to a derivative with
respect to the explicit thermal dependence, not the implicit one,
hidden in $\bar G.$ \rt{, see Appendix~\ref{app:id}.} As before, this can be rewritten as:
\beq\label{eq:TflowM}
\frac{\partial \bar M^2}{\partial T}=\frac{\bar V}{2}\left(\frac{\partial}{\partial T}\int_{Q<\Lambda}^T\right)\bar G\,.
\eeq
A similar strategy can be used to express the infinitesimal variations of $\bar V$ itself. We obtain:
\beq
\frac{\partial\bar V}{\partial \phi^2} & = & \frac{\bar V^3}{2}\int_{Q<\Lambda}^T\bar G^3\,,\label{eq:flowVphi}\\
\frac{\partial\bar V}{\partial T} & = & -\frac{\bar V^2}{2}\left(\frac{\partial}{\partial T}\int_{Q<\Lambda}^T\right)\bar G^2+\frac{\bar V^3}{2}\left(\int_{Q<\Lambda}^T\bar G^3\right)\left(\frac{\partial}{\partial T}\int_{Q<\Lambda}^T\right)\bar G\label{eq:TflowV}\,.
\eeq
The interesting feature of Eqs.~(\ref{eq:flowMphi}), (\ref{eq:TflowM}), (\ref{eq:flowVphi}) and (\ref{eq:TflowV}) is that they are explicitly finite, without the need to impose any renormalization condition. They can be used as an alternative tool to solve the gap equation. The only information that remains to be specified are the initial conditions for $\bar M^2$ and $\bar V$, for instance at $\phi^2=0$ and $T=T_\star$. The initial condition for $\bar M^2$ reads $\bar M^2_{\phi=0,\,T_\star}=m^2_\star$ which is nothing but the renormalization condition (\ref{eq:rencondm}). As for $\bar V$, its initial condition reads:
\beq\label{eq:rencondV}
\bar V_{\phi=0,T_\star}=\lambda_\star\,,
\eeq
and it is easily checked that it can be reinterpreted as the the renormalization condition corresponding to the choice of $\lambda_0$ in Eq.~(\ref{eq:l0}). Thus, specifying the initial conditions in the flow approach corresponds to imposing renormalization conditions in the standard approach\footnote{This is similar in spirit to the approach followed in \cite{Blaizot:2010zx}.}. As already discussed, $m^2_\star$ and $\lambda_\star$ are taken strictly positive.\\
In practice, it is more convenient to solve the evolution equations (\ref{eq:flowMphi}) and (\ref{eq:TflowM}) coupled to the explicit and renormalized expression for $\bar V$ which we obtain from combining Eqs.~(\ref{eq:BS}) and (\ref{eq:l0}):
\beq\label{eq:BSren}
\frac{1}{\bar V}=\frac{1}{\lambda_\star}+\frac{1}{2}\left[\int_{Q<\Lambda}^T\bar G^2-\int_{Q<\Lambda}^{T_\star} G_\star^2\right]\,.
\eeq
Because $\phi^2_{\rm c}(T_\star)<0$, we first solve the equation (\ref{eq:flowMphi}) at $T=T_\star$ from the initial condition $\bar M^2_{\phi=0,\,T_\star}=m_\star^2>0$ and obtain the function $\bar M^2_{\phi,\,T_\star}$. For each value of $\phi$, the solution $\bar M^2_{\phi,\,T_\star}$ is then used as the initial condition for the flow equation (\ref{eq:TflowM}), which we solve from $T_\star$ for decreasing values of $T$. We obtain in this way the function $\bar M^2_{\phi,\,T}$ for any value of $\phi$ and $T$ (with $\phi^2\geq\phi^2_{\rm c}(T)$). This method will be used in the next section where we evaluate the effective potential. Notice finally that $\bar V=-\lambda_\star/g'_\Lambda(\bar M^2)$ where
$g_\Lambda(M^2)$ is the function introduced in Eq.~(\ref{eq:gapr}) to
discuss the solutions of the renormalized gap equation (see
Eq.~(\ref{eq:deux}) for its derivative). We are only interested in the physical solution to the gap equation. Irrespectively of the value of $\Lambda$, this solution is such that $g'_\Lambda(\bar M^2)<0$, see Appendix~\ref{app:sol}, and thus $\bar V>0$. From Eqs.~(\ref{eq:flowMphi}) and (\ref{eq:flowVphi}), it follows then that $\bar M^2$ and $\bar V$ increase with $\phi^2$. In particular at $T=T_\star$ and for any $\phi^2$, one has $\bar M^2\geq m^2_\star$ and $\bar V\geq\lambda_\star$. We will make use of this remark later.
\section{Effective potential}\label{sec:eff_pot}
Once the mass $\bar M$ and the propagator $\bar G$ have been
determined, one can evaluate numerically the renormalized version of the
effective potential to be given in Section \ref{sec:ren_eff_pot} and study its change of shape as the temperature
$T$ is lowered from the ``high'' temperature $T_\star$, possibly
triggering a change of phase. In this section we show that a large
part of the analysis can be done analytically. We prove in particular
that in the Hartree approximation, for those values of the parameters such that there is a phase transition,
the later cannot be of second order. To this purpose, we use information on
the geometry of the effective potential, such as its curvature, as
well as information on the location of its possible extrema, as given
by the field equation. Notice that our analysis will be performed in the presence of a cut-off $\Lambda$ below the Landau scale $\Lambda_{\rm p}$. After renormalization of the effective potential, see below, and for parameters such that $\Lambda_{\rm p}$ is much larger than all other scales in the problem, the renormalized effective potential computed with $\Lambda<\Lambda_{\rm p}$ will be essentially the same as the one computed in the limit $\Lambda\rightarrow\infty$. We shall use this simplification when evaluating the effective potential numerically.
\subsection{Geometry of the effective potential}\label{sec:geom}
The geometry of the effective potential is encoded in its field
derivatives. For instance, the curvature for a given value of $\phi$
reads:
\beq\label{eq:g2}
\hat M^2\equiv\frac{\delta^2\gamma}{\delta\phi^2}=m_0^2+\frac{\lambda_0}{2}\,\phi^2
+\frac{\lambda_0}{2}\int_{Q<\Lambda}^T\bar G -\frac{\lambda_0}{2}\,\phi\int_{Q<\Lambda}^T\bar G^2\,\frac{\partial\bar M^2}{\partial\phi}\,,
\eeq
where we made use of the stationnarity condition (\ref{eq:stat}) and $\delta \bar G=-\bar G^2\delta \bar G^{-1}$. From
the flow equation (\ref{eq:flowMphi}), one obtains $\partial\bar M^2/\partial\phi=\bar V\phi$. Plugging this into Eq.~(\ref{eq:g2}) and using Eq.~(\ref{eq:BS}), we arrive at:
\beq\label{eq:g2_bis}
\hat M^2=\bar M^2+\bar V\,\phi^2-\lambda_0\,\phi^2\,.
\eeq
Similarly, the fourth derivative of the effective potential at $\phi=0$ is given by:
\beq\label{eq:g4}
\hat V_{\phi=0}\equiv\left.\frac{\delta^4\gamma}{\delta\phi^4}\right|_{\phi=0}
=\lambda_0-\frac{3}{2}\,\lambda_0\int_{Q<\Lambda}^T\bar G^2_{\phi=0}\left.\frac{\partial^2\bar M^2}{\partial\phi^2}\right|_{\phi=0}\,.
\eeq
The same argument as for $\hat M^2$ leads to:
\beq\label{eq:g4_bis}
\hat V_{\phi=0}=3\bar V_{\phi=0}-2\lambda_0\,.
\eeq
These results illustrate the general discussion given in the Introduction concerning the existence of different definitions of $n$-point functions in a given approximation of the 2PI effective action. For instance, the two-point function (here simply a mass) could be defined either from the solution $\bar M^2$ of the gap equation or from the second derivative $\hat M^2$ of the effective potential. Although these two definitions should coincide for any value of $\phi$ and $T$ in the absence of approximations, Eq.~(\ref{eq:g2_bis}) clearly shows that, in the Hartree approximation, they coincide for $\phi=0$ only. For other approximations, the discrepancy between the different definitions of the two-point function can also be observed for $\phi=0$. Similar remarks apply to the four-point function. Notice however that these discrepancies always appear as higher order effects. In the present case for instance, if we consider a formal expansion in power of the bare coupling $\lambda_0$, we obtain $\hat M^2-\bar M^2=\mathcal{O}(\lambda_0^2)$ as well as $\hat{V}-\bar V={\cal O}(\lambda_0^2)$ which show that the discrepancies are beyond the accuracy of the Hartree approximation.\\
More generally the discrepancies between the different definitions of a given $n$-point function are always of higher order as compared to the accuracy of the 2PI truncation that one considers and are thus of little relevance a priori. However, this is only true for $n\geq 6$. For $n=2$ and $n=4$, the discrepancies between the different definitions of $n$-point functions can be divergent, and they are thus negligeable only after divergences have been properly absorbed. For instance, the two different definitions of the two-point function can bring independent quadratic divergences. In order to remove them, one needs to allow for two independent bare masses following the general prescription of \cite{Berges:2005hc}. It turns out that in the case of the Hartree approximation, the quadratic divergences are not independent but equal\footnote{This has to do with the fact that the two definitions of the mass coincide for $\phi=0$.} which allows one to work with only one bare mass. Nevertheless the sensitivities $\bar M^2$ and $\hat M^2$ with respect to $\Lambda$ are radically different for, as we have seen above, $\bar M^2$ is defined for any value of $\Lambda$ with $\Lambda$-dependent terms of the form $1/\Lambda^2$, whereas, due to the presence of $\lambda_0$ in Eq.~(\ref{eq:g2_bis}), $\hat M^2$ diverges as $\Lambda\rightarrow\Lambda_{\rm p}^-$. For the same reason, the sensitivities of $\bar V$ and $\hat V$ with respect to $\Lambda$ are different. One possible way out is to argue, in a way similar to \cite{AmelinoCamelia:1992nc}, that if one considers $\Lambda\ll\Lambda_{\rm p}$, but much larger than all other scales in the problem, the additional sensitivity of $\hat M^2$ and $\hat V$ as compared to $\bar M^2$ and $\bar V$ is a very mild one. Here we follow a different approach,\footnote{Despite this different point of view, our conclusions regarding the order of the phase transition will be the same as those in \cite{AmelinoCamelia:1992nc}.} in line with the general presentation in \cite{Berges:2005hc}, which is generalizable to higher order approximations. Since the difference of sensitivity with respect to $\Lambda$ comes from the fact that the two definitions $\bar V$ and $\hat V$ of the four-point function differ within a given approximation, it is natural to introduce a bare coupling $\lambda_4$ different from $\lambda_0$ in such a way that the original 2PI functional reads
\beq
\gamma[\phi,G] & = & \frac{m_0^2}{2}\,\phi^2+\frac{\lambda_4}{4!}\,\phi^4+\frac{1}{2}\int_{Q<\Lambda}^T\ln G^{-1}+\frac{1}{2}\int_{Q<\Lambda}^T(Q^2+m_0^2)\,G\nonumber\\
& + & \frac{\lambda_0}{4}\,\phi^2\int^T_{Q<\Lambda} G+\frac{\lambda_0}{8}\left(\int_{Q<\Lambda}^T G\right)^2.
\label{eq:gen_new}
\eeq
Clearly $\lambda_0$ enters the definition of $\bar V$, as before,
whereas $\lambda_4$ appears as a tree level contribution to $\hat
V$. With this modification, using the same steps as those leading to
Eqs.~(\ref{eq:g2_bis}) and (\ref{eq:g4_bis}) we obtain
\beq\label{eq:Mhat_w_l4}
\hat M^2 & = & \bar M^2+\bar V\,\phi^2+\frac{1}{2}(\lambda_4-3\lambda_0)\phi^2\,,\\
\hat V_{\phi=0} & = & 3\bar V_{\phi=0}+(\lambda_4-3\lambda_0)\,.
\eeq
We can now adjust $\lambda_4$ such that $\hat V$ is as insensitive to $\Lambda$ as $\bar V$, that is up to terms of order $1/\Lambda^2$. This however does not fix the finite part of $\lambda_4$. In order to fix the later without introducing additional parameters, and following \cite{Berges:2005hc,Reinosa:2009tc}, we impose the consistency condition:
\beq\label{eq:conscond}
\hat V_{\phi=0,\,T_\star}=\bar V_{\phi=0,\,T_\star}\,,
\eeq
at the renormalization point $\phi=0$ and $T=T_\star$, condition which should hold automatically if no approximation was considered. This condition translates into a particular choice of the bare parameter $\lambda_4$:
\beq\label{eq:l4}
\lambda_4-3\lambda_0=-2\lambda_\star\,.
\eeq
Using this particular choice, the derivatives of the effective potential become then
\beq
\hat M^2 & = & \bar M^2+(\bar V-\lambda_\star)\phi^2\,,\label{eq:curvren}\\
\hat V_{\phi=0} & = & 3\bar V_{\phi=0}-2\lambda_\star\,.
\eeq
Notice that $\hat M^2$ and $\hat V$ are now both cut-off insensitive: the renormalization of $\hat V$ automatically took into account the renormalization of $\hat M^2$. We stress that this is only specific to certain approximations. In general, one would need to introduce a second bare mass which one would fix by imposing another consistency condition. Notice also that, if we expand the bare couplings in powers of the renormalized coupling $\lambda_\star$ using Eqs.~(\ref{eq:l0}) and (\ref{eq:l4}), we find that $\lambda_4-\lambda_0=\mathcal{O}(\lambda^2_\star)$, in line with the fact that the artifacts that the introduction of a second bare coupling $\lambda_4$ is meant to cure lie beyond the accuracy of the Hartree approximation. As a final remark, recall that, as we have shown in the previous section, $\bar M^2\geq m^2_\star>0$ and $\bar V\geq \lambda_\star>0$ at $T=T_\star$ and for any $\phi^2$. It follows that $\hat M^2\geq m_\star^2>0$ at $T=T_\star$ and for any $\phi^2$. This means that the effective potential at the initial temperature $T=T_\star$ is strictly convex.
\subsection{Renormalized effective potential}\label{sec:ren_eff_pot}
We are now in a position to derive a finite expression for the effective potential. Using
Eq.~(\ref{eq:gap}) in Eq.~(\ref{eq:gen_new}) the effective potential
$\gamma(\phi)\equiv\gamma[\phi,\bar G]$ reads
\beq
\gamma(\phi)=\frac{1}{4!}(\lambda_4-3\lambda_0)\phi^4+\frac{\bar M^4-m_0^4}{2\lambda_0}+\frac{1}{2}\int_{Q<\Lambda}^T\ln \bar G^{-1}+\frac{1}{2}\int_{Q<\Lambda}^TQ^2\,\bar G\,.
\eeq
Using Eqs.~(\ref{eq:BS}) and (\ref{eq:l4}), we arrive at
\beq
\gamma(\phi)=-\frac{m_0^4}{2\lambda_0}-\frac{\lambda_\star}{12}\,\phi^4+\frac{\bar M^4}{2\bar V}+\frac{1}{2}\int_{Q<\Lambda}^T\left[\ln \bar G^{-1}+Q^2\,\bar G-\frac{1}{2}\,\bar M^4\,\bar G^2\right]\,.
\eeq
Expanding the integrand for large $Q$ we observe that it has no
quadratic or logarithmic divergence, only a quartic one of the form
$\int_{Q<\Lambda}^T(\ln Q^2+1)$. We can thus write
\beq
\gamma(\phi)=-\frac{m_0^4}{2\lambda_0}-\frac{\lambda_\star}{12}\,\phi^4+\frac{\bar M^4}{2\bar V}+\frac{1}{2}\int_{Q<\Lambda}^T\left[\ln (Q^2+\bar M^2)-\ln Q^2-\bar M^2\,\bar G-\frac{1}{2}\,\bar M^4\,\bar G^2\right]+\frac{1}{2}\int_{Q<\Lambda}^T\Big[\ln Q^2+1\Big]\,.
\eeq
The first integral is convergent. The second is divergent, but one can convince oneself that its
divergence is $T$- and $\phi$-independent. Then, if we subtract the
effective potential evaluated at $T_\star$ and $\phi=0$,
we obtain an explicitly convergent expression
\beq
\label{eq:gamma_final}
\gamma(\phi)-\gamma_{\star}(0)=-\frac{\lambda_\star}{12}\,\phi^4+\frac{\bar M^4}{2\bar V}-\frac{m_\star^4}{2\lambda_\star} & + & \frac{1}{2}\int_{q<\Lambda}\left[\varepsilon_q-q-\frac{\bar M^2}{2\varepsilon_q}-\frac{\bar M^4}{8\varepsilon_q^3}\right]- \frac{1}{2}\int_{q<\Lambda}\left[\varepsilon^\star_q-q-\frac{m_\star^2}{2\varepsilon^\star_q}-\frac{m_\star^4}{8{\varepsilon^\star_q}^3}\right]\nonumber\\
& - & \frac{1}{2}\int_{q<\Lambda} \left[-2T\ln(1-e^{-\varepsilon_q/T})+\bar M^2\,\frac{n_{\varepsilon_q}}{\varepsilon_q}+\bar M^4\,\frac{n_{\varepsilon_q}-\varepsilon_q\,n'_{\varepsilon_q}}{4\varepsilon_q^3}\right]\nonumber\\
& + & \frac{1}{2}\int_{q<\Lambda} \left[-2T_\star\ln(1-e^{-\varepsilon^\star_q/T_\star})+m_\star^2\,\frac{n^\star_{\varepsilon^\star_q}}{\varepsilon^\star_q}+m_\star^4\,\frac{n^\star_{\varepsilon^\star_q}-\varepsilon^\star_q\,{n^\star}'_{\varepsilon^\star_q}}{4{\varepsilon^\star_q}^3}\right]\,,
\eeq
which is our final form for the renormalized effective potential.\footnote{The first two integrals can be done analytically with the
result $(\bar M^4-m_\star^4)/(128\pi^2)$ in the limit $\Lambda\to\infty$.} This formula for the effective potential differs from that in \cite{AmelinoCamelia:1992nc} in that we have used the additional bare parameter $\lambda_4$. The approach we have followed here to obtain a finite effective potential can be generalized to higher orders of approximation, see for instance \cite{Arrizabalaga:2006hj}.
\subsection{Extrema of the effective potential}\label{sec:extrema}
At any temperature, the extrema of the effective potential (\ref{eq:gen_new}) are given by the field equation:
\beq
0=\left.\frac{\delta\gamma}{\delta\phi}\right|_{\bar\phi}\,.
\eeq
In making this equation explicit one can take advantage of the stationarity condition (\ref{eq:stat}). In the Hartree approximation, and taking into account the newly introduced bare coupling $\lambda_4$, we arrive at:
\beq\label{eq:field}
0=\bar\phi\left(m_0^2+\frac{\lambda_4}{6}\,\bar\phi^2+\frac{\lambda_0}{2}\int_{Q<\Lambda}^T\bar G_{\bar\phi}\right)\,.
\eeq
Using the relation (\ref{eq:l4}) as well as the gap equation (\ref{eq:gap}), we
finally arrive at the system of equations:
\beq\label{eq:renfield}
0=\bar\phi\left(\frac{\lambda_\star}{3}\bar\phi^2-\bar M^2_{\bar\phi}\right) \quad {\rm and} \quad 0=g_\Lambda(\bar M^2_{\bar\phi},\bar\phi^2)\,,
\eeq
where we have made explicit the dependence of $g_\Lambda(M^2)$ with respect to $\phi$. Notice that this system of equations is explicitly finite. This is not surprising for, once the second and fourth derivatives of the effective potential have been renormalized, the effective potential is renormalized (up to an overall constant), and so is any information that can be extracted from it such as the location of its extrema.\footnote{Notice that the way we have obtained an explicitly renormalized field equation differs in spirit from that followed in \cite{Lenaghan:1999si} for the case of the $O(N)$ model. There, the field equation was similar to the one in Eq.~(\ref{eq:renfield}) with $\lambda_\star$ replaced by the bare coupling (denoted by $\lambda$ in Ref.~\cite{Lenaghan:1999si}). Since the latter had no divergence to absorb in the field equation, it was then identified to the renormalized coupling (although this identification was not possible at the level of the gap equation). Even though this result is correct and agrees with Eq.~(\ref{eq:renfield}), its justification in \cite{Lenaghan:1999si} hides the origin of the difficulty, namely the existence of multiple definitions for $n$-point functions and does not allow a generalization to higher orders of approximation. For a discussion of similar issues, see also \cite{Destri:2005qm}.}\\
According to Eq.~(\ref{eq:renfield}), the trivial extremum $\bar\phi=0$ only makes sense as long as the gap equation at $\phi=0$ has a solution. This ceases to be true if $\phi^2_{\rm c}(T)$ is strictly positive, that is when $C_\star<0$ and $T<T_{\rm c}$, see our earlier discussion. On the other hand, the non trivial extrema obey the system:
\beq
0=\frac{\lambda_\star}{3}\bar\phi^2-\bar M^2_{\bar\phi} \quad {\rm and} \quad 0=g_\Lambda(\bar M^2_{\bar\phi},\bar\phi^2)\,,
\eeq
which we rewrite conveniently as
\beq
\bar\phi^2=\frac{3}{\lambda_\star}\bar M^2_{\bar\phi} \quad {\rm and} \quad 0=h_\Lambda(\bar M^2_{\bar\phi})\,,
\eeq
where the function $h_\Lambda(M^2)$ does not depend on the field and is given by
\beq\label{eq:h}
h_\Lambda(M^2)\equiv g_\Lambda\left(M^2,\frac{3}{\lambda_\star}M^2\right)=\frac{3}{2}M^2+g_\Lambda(M^2,\phi^2=0)\,,
\eeq
with first and second derivatives equal to
\beq
h'_\Lambda(M^2) & = & \frac{3}{2}+g'_\Lambda(M^2)\,,\\
h''_\Lambda(M^2) & = & g''_\Lambda(M^2)\,.
\eeq
The non trivial extrema of the effective potential can thus be obtained from the zeros of $h_\Lambda(M^2)$, which we discuss below. Notice that, from Eq.~(\ref{eq:curvren}), the curvature of the effective potential at the non-trivial extrema is given by
\beq\label{eq:curvextrema}
\hat M^2=\left(\bar V_{\bar\phi}-\frac{2\lambda_\star}{3}\right)\bar\phi^2=-\lambda_\star\left(\frac{1}{g'_\Lambda(\bar M^2_{\bar\phi})}+\frac{2}{3}\right)\bar\phi^2=-\frac{2\lambda_\star}{3}\frac{h'_\Lambda(\bar M^2_{\bar\phi})}{g'_\Lambda(\bar M^2_{\bar\phi})}\,\bar\phi^2\,.
\eeq
Because $g'_\Lambda(\bar M^2)<0$ when $\Lambda<\Lambda_{\rm p}$, see Appendix~\ref{app:sol}, the sign of the curvature is the sign of $h'_\Lambda(\bar M^2)$.
\subsection{Temperature dependence of the effective potential}\label{sec:temp}
Let us now discuss the zeros of $h_\Lambda(M^2)$ and the corresponding curvature encoded in Eq.~(\ref{eq:curvextrema}) and deduce the shape of the effective potential as we decrease the temperature from $T_\star$ down to $T=0$. Because $h''_\Lambda(M^2)>0$, the function $h'_\Lambda(M^2)$ increases strictly from $h'_\Lambda(0)=-\infty$ to
\beq
h'_\Lambda(\infty)=\frac{3}{2}+g'_\Lambda(\infty)=\frac{3}{2}-\frac{\lambda_\star}{\lambda_0}=\frac{1}{2}+\frac{\lambda_\star}{2}\int_{Q<\Lambda}^{T_\star}G_\star^2>0\,,
\eeq
where we have used $g'_\Lambda(\infty)=-\lambda_\star/\lambda_0$, see Appendix~\ref{app:sol}, and Eq.~(\ref{eq:l0}). The function $h'_\Lambda(M^2)$ has to vanish for some $\bar M^2_{\rm e}$. It follows that the function $h_\Lambda(M^2)$ has a minimum at $M^2=\bar M^2_{\rm e}$: $h_\Lambda(M^2)$ decreases strictly from $h_\Lambda(0)$ to $h_\Lambda(\bar M^2_{\rm e})$ and then increases towards $h_\Lambda(\infty)=\infty$. The number of non-trivial extrema of the effective potential depends thus on the signs of $h_\Lambda(0)$ and $h_\Lambda(\bar M^2_{\rm e})$ which we now discuss.\\
Let us consider $h_\Lambda(0)$ first. It is nothing but $g_\Lambda(0)$ at $\phi=0$, see Eq.~(\ref{eq:h}), and thus from the discussion above Eq.~(\ref{eq:appp}):
\beq
h_\Lambda(0)=-\frac{\lambda_\star}{2}\,\phi^2_{\rm c}(T,\Lambda)=C_\star+\frac{\lambda_\star}{2}\int_{q<\Lambda}\frac{n_q}{q}\,,
\eeq
where $C_\star$ was defined in Eq.~(\ref{eq:Cstar}) and depends on the
parameters. As we have seen below Eq.~(\ref{eq:Tc}), $\phi^2_{\rm c}(T,\Lambda)$ is strictly negative for $T=T_\star$. It follows that $h_\Lambda(0)$ is strictly positive for $T=T_\star$. Moreover, the previous formula shows that
$h_\Lambda(0)$ decreases strictly as we lower $T$ and reaches
$C_\star$ at $T=0$. If the parameters are such that $C_\star>0$, then
$h_\Lambda(0)$ remains strictly positive down to $T=0$. If
$C_\star=0$, $h_\Lambda(0)$ remains strictly positive down to $T>0$
and vanishes at $T=0$. If $C_\star<0$, $h_\Lambda(0)$ remains strictly
positive while $T>T_{\rm c}$ ($0<T_{\rm c}<T_\star$), vanishes at
$T=T_{\rm c},$ changes sign and remains strictly negative down to $T=0$. The same type of analysis needs to be done for $h_\Lambda(\bar M^2_{\rm e})$. Using Eq.~(\ref{eq:gapr}), we obtain
\beq
h_\Lambda(\bar M^2_{\rm e})=\frac{1}{2}\,\bar M^2_{\rm e}+m^2_\star+\frac{\lambda_\star}{2}\left[\int_{q<\Lambda}\frac{\delta_\star n_{\varepsilon^{\rm e}_q}}{\varepsilon^{\rm e}_q}+(\bar M^2_{\rm e}-m^2_\star)^2\int^{T_\star}_{Q<\Lambda} G_\star^2\,\bar G_{\rm e}\right].
\eeq
Clearly, $h_\Lambda(\bar M_{\rm e})>0$ at $T=T_\star$. Moreover,
because $h'_\Lambda(\bar M_{\rm e})=0$, the thermal dependence of
$h(\bar M_{\rm e})$ is the explicit one, encoded in the tadpole
integral. It follows that $h_\Lambda(\bar M^2_{\rm e})$ decreases as
one decreases $T$. We shall denote by $D_\star$ its value at $T=0$. If
the parameters are such that $D_\star>0$, $h_\Lambda(\bar M^2_{\rm
e})$ remains strictly positive all the way down to $T=0$. If
$D_\star=0$, it remains strictly positive down to $T>0$ and vanishes
at $T=0$. If $D_\star<0$, $h_\Lambda(\bar M^2_{\rm e})$ remains
strictly positive while $T>T_{\rm s}$ ($0<T_{\rm s}<T_\star$),
vanishes at $T=T_{\rm s},$ changes sign and remains strictly negative down to $T=0.$\\
A priori, discussing the signs of $h_\Lambda(0)$ and $h_\Lambda(\bar M^2_{\rm e})$ involves nine cases, depending on the values of $C_\star$ and $D_\star$, and thus nine regions in parameter space. However, because $C_\star$ and $D_\star$ are the values of $h_\Lambda(0)$ and $h_\Lambda(\bar M^2_{\rm e})$ at $T=0$ and $h_\Lambda(0)>h_\Lambda(\bar M^2_{\rm e})$, we have necessarily $C_\star>D_\star$, which reduces the number of regions to five:\\
\begin{figure}[!tbp]
\begin{center}
\includegraphics[keepaspectratio,width=0.9\textwidth,angle=0]{./eff_pots}
\caption{The temperature evolution of the effective potential obtained for $m^2_\star/T_\star^2=0.1$, $\lambda_\star=3$ in the limit $\Lambda\rightarrow\infty$ by evaluating Eq.~(\ref{eq:gamma_final}) with $\bar M^2$ calculated
at $T_\star$ using the $\phi^2$-flow Eq.~(\ref{eq:flowMphi}) and at $T\ne0$
using the $T$-flow Eq.~(\ref{eq:TflowM}). On each plot, a constant has been subtracted for convenience, without affecting the physical interpretation.
\label{fig:eff_pot}}
\end{center}
\end{figure}
1. Region corresponding to $C_\star<0$: For
such parameters, $h_\Lambda(\bar M^2_{\rm e})$ changes sign at
$T_{\rm s}$ and $h_\Lambda(0)$ changes sign at $T_{\rm c}<T_{\rm
s}$. The evolution of the effective potential as we decrease $T$
from $T_\star$ down to $T=0$ goes as follows:
\begin{itemize}
\item[$-$] For $T_{\rm s}<T\leq T_\star$, both $h_\Lambda(0)$ and $h_\Lambda(\bar M^2_{\rm e})$ are strictly positive and the effective potential has only a trivial extremum at $\phi=0$. It is a minimum because $\hat M^2=\bar M^2>0$ at $\phi=0$. The effective potential is convex, see Fig.~\ref{fig:eff_pot}(a).
\item[$-$] At $T=T_{\rm s}$, $h_\Lambda(0)$ is still strictly positive but $h(\bar M^2_{\rm e})$ vanishes: in addition to the trivial minimum at $\phi=0$, a non trivial extremum appears. According to Eq.~(\ref{eq:curvextrema}), its curvature is zero. The only possibility is that it is an inflection point of the effective potential, see Fig.~\ref{fig:eff_pot}(b).
\item[$-$] For $T_{\rm c}<T<T_{\rm s}$, $h_\Lambda(0)$ is still
strictly positive and $h(\bar M^2_{\rm e})$ is strictly negative: in
addition to the trivial minimum at $\phi=0$, two non-trivial extrema
are present. According to Eq.~(\ref{eq:curvextrema}), the
non-trivial extremum closest to $\phi=0$ is a maximum and the other
one is a minimum, see Fig.~\ref{fig:eff_pot}(c). As it will become
clear in the next point, it turns out that in this temperature interval
there is a temperature $T_{\rm t}$ ($T_{\rm c}<T_{\rm t}<T_{\rm s}$)
at which the minima are degenerate and below which the non trivial minimum
becomes the absolute minimum, see Fig.~\ref{fig:eff_pot}(d).
\item[$-$] At $T=T_{\rm c}$, $h_\Lambda(0)$ vanishes and
$h_\Lambda(\bar M^2_{\rm e})$ is strictly negative: the non-trivial
extremum which corresponded to a maximum is at $\phi=0$ and merges
with the trivial minimum. The curvature at $\phi=0$ is zero. Because
there is still a non-trivial minimum at $\phi\neq 0$, the only
possibility is that at $T=T_{\rm c}$ the effective potential
presents a maximum at $\phi=0$, see Fig.~\ref{fig:eff_pot}(e), with
the consequence that there was a temperature $T_{\rm t}$ ($T_{\rm c}<T_{\rm t}<T_{\rm s}$) below which the non-trivial minimum became the absolute minimum.
\item[$-$] For $T<T_{\rm c}$, the gap equation has no solution at $\phi=0$: the trivial extremum disappears. The effective potential is defined only for $\phi^2\geq\phi^2_{\rm c}(T)$ and has a single non-trivial minimum, see Fig.~\ref{fig:eff_pot}(f).
\end{itemize}
2. Region corresponding to $C_\star=0$: It is just a line in parameter space. The same analysis as above applies except from the fact that $T_{\rm c}=0$. The final shape of the potential is that of Fig.~\ref{fig:eff_pot}(e).\\
3. Region corresponding to $C_\star>0$ and $D_\star<0$: The same
analysis applies except from the fact that there is no $T_{\rm
c}$. The final shape of the potential at $T=0$ is either that of Fig.~\ref{fig:eff_pot}(d) or that of Fig.~\ref{fig:eff_pot}(c), but without evaluating the potential we cannot decide whether or not the non-trivial minimum became the absolute one at some temperature.\\
4. Region corresponding to $C_\star>0$ and $D_\star=0$: It is just a line in parameter space. The same analysis applies except from the fact that there is no $T_{\rm c}$ and $T_{\rm s}=0$. The final shape of the potential is that of Fig.~\ref{fig:eff_pot}(b).\\
5. Region corresponding to $C_\star>0$ and $D_\star>0$: There are no
$T_{\rm c}$ or $T_{\rm s}$. The final shape of the potential is
convex, see Fig.~\ref{fig:eff_pot}(a).\\
The previous discussion shows that, if there is a phase transition, it is first order due to the finite jump in the location of the absolute minimum of the effective potential (see Fig.~\ref{fig:eff_pot}). The transition occurs at $T=T_{\rm t}$ while $T_{\rm c}$ and $T_{\rm s}$ correspond to the upper and lower spinodal temperatures at which a maximum and a minimum of the potential merge.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics[keepaspectratio,width=0.6\textwidth,angle=0]{./params}
\caption{Parameter space and order of the phase transition. The labels
on the curves indicate the value of $\ln(\Lambda_{\rm p}/m_\star)$. We focus on the region $\ln(\Lambda_{\rm p}/m_\star)>5$ which is such our results can be considered cut-off insensitive for a cut-off $\Lambda$ below the Landau scale. Note that this region lies below the dot-dashed curve that is in the region of validity of the condition (\ref{eq:cond_on_RoI}) which played a role in part of our analysis of the gap equation.
\label{fig:params}}
\end{center}
\end{figure}
These conclusions are valid for a given value of the cut-off $\Lambda<\Lambda_{\rm p}$. As we have explained in the previous section, if we want the results to be independent of the cut-off $\Lambda$ we need to choose the parameters such that $\Lambda_{\rm p}\gg m_\star,\,T_\star,\,T$ so that it is possible to choose $\Lambda$ below the Landau scale but larger with respect to all other scales.\footnote{For practical purposes, it is then enough to compute the effective potential for $\Lambda=\infty$ without introducing any substantial difference as compared to the effective potential at $\Lambda<\Lambda_{\rm p}$. One can also perform an analytical study of the effective potential for $\Lambda>\Lambda_{\rm p}$. The only subtlety is that $g'(M^2)$ can be positive for $M^2>\bar M^2_{\rm p}$ and the discussion around Eq.~(\ref{eq:curvextrema}) needs to be revisited. One can convince oneself that the zeros of $h(M^2)$ are all low scales, lying below the high scale $\bar M^2_{\rm p}$, and correspond thus to negative values of $g'(M^2)$.} In what follows we choose $T_\star=1$ which sets the mass unit. Since we are only interested in values of $T$ below $T_\star$, it is then enough to choose $m_\star$ and $\lambda_\star$ such that $\Lambda_{\rm p}/m_\star\gg 1$. For definiteness, we choose to work in the region of parameter space such that $\ln(\Lambda_{\rm p}/m_\star)>5$. Note that, in this regime
\beq
\frac{\bar M^2_{\rm e}(T=0)}{m^2_\star}=\frac{m_\star e}{2\Lambda_{\rm p}}\exp\left(-24\pi^2\int_q\frac{n^\star_{\varepsilon^\star_q}-\varepsilon^\star_q {n^\star}'_{\varepsilon^\star_q}}{2{\varepsilon^\star_q}^3}\right)\ll 1\,,
\eeq
which we obtained from the equation $h'_\Lambda(\bar M^2_{\rm e})=0$ at $T=0$ and for $\Lambda\gg m_\star,\,T_\star$. One can show in the same way that $D_\star$, which is defined as $h_\Lambda(\bar M^2_{\rm e})$ at $T=0$, is given by
\beq
D_\star=C_\star-\frac{\lambda_\star}{32\pi^2}\bar M^2_{\rm e}(T=0)\,.
\eeq
Since regions 2., 3. and 4. are such that $C_\star\geq 0$ and $D_\star\leq 0$, it follows that $0\leq C_\star\leq\lambda_\star \bar M^2_{\rm e}(T=0)/(32\pi^2)$ and they form a very narrow band, indistinguishable from the line $C_\star=0$ whose equation is\footnote{For practical purposes this curve can be approximated using the high temperature expansion: $\lambda_\star(m_\star)=2m_\star^2/(T_\star^2/12-m_\star T_\star/(8\pi))$.}
\beq\label{eq:Cstar_HTE}
\lambda_\star(m_\star)=2m^2_\star\left[\int_{q}\frac{n^\star_q}{q}-m^4_\star \int^{T_\star}_{Q} \frac{G_\star^2}{Q^2}\right]^{-1}.
\eeq
The parameter space is then essentially divided in two regions, see Fig.~\ref{fig:params}. The lower (grey) region, corresponding to region 5. of the analysis above, is such that no phase transition occurs between $T=T_\star$ and $T=0$. The upper (white) region, corresponding to region 1. of the analysis above, is such that a first order phase transition occurs at $T=T_{\rm t}$. The temperatures $T_{\rm t}$ and $T_{\rm s}$ can only be accessed numerically. We find that for moderate values of the parameters
$\lambda_\star$ and $m^2_\star$ the first order phase transition is
rather weak, that is for reasonable values of the parameters the
difference between $T_{\rm c}$ and $T_{\rm s}$ is small and also for $T_{\rm c}<T<T_{\rm s}$
the non-trivial minimum is close to the trivial one.
The first order nature of the phase
transition is an artifact of the Hartree approximation which corresponds to a $\mathcal{O}(\lambda)$ truncation of the 2PI functional at skeleton level. Since this effect is rather weak, an improvement of the truncation by including the order-$\lambda^2$ and field-dependent skeleton contribution to the 2PI functional seems to turn the phase transition into second order \cite{Arrizabalaga:2006hj}, in accordance with universality arguments.\\
For phenomenological applications a parametrization of the model is
needed, which is usually done at zero temperature. For this reason it
is interesting to know which values of the parameters $m^2_\star$ and $\lambda_\star$
at $T_\star$ are of interest. In the region of the parameter space
where there is a phase transition $T_c$ increases steeply with
$\lambda_\star.$ Therefore, in order for the phase transition to occur at
a reasonable value of temperature compared to $T_\star$ one needs to
restrict the possible values of the parameters in the
$m^2_\star-\lambda_\star$ space. The accessible values of
$\bar\phi(T=0)$ and $\bar M(T=0)$ when the restriction
$T_c/T_\star \le 0.3$ is imposed can be seen
Fig.~(\ref{fig:params_cut}). This means that the region of the
parameter space which is interesting in practice is a relatively
narrow one, close to the curve given in Eq.~(\ref{eq:Cstar_HTE}).
Setting $T_\star=1$~GeV, these values are in the range of the pion
decay constant, pion masses, and the transition temperature in the
$O(4)$ linear sigma model.
\begin{figure}[htbp]
\begin{center}
\includegraphics[keepaspectratio,width=0.55\textwidth,angle=0]{./params_cut}
\caption{ Dependence of the solution of the field equation at $T=0$ on
the parameters in the region of the parameter space where
$T_{\rm c}/T_\star \le 0.3$ ($T_c$ increases with $\lambda_\star.$)
The labels near the solid curves refers to the solution $\bar M$ of the gap
equation at $T=0.$
\label{fig:params_cut}}
\end{center}
\end{figure}
\pagebreak
\section{Conclusions and discussion\label{sec:concl}}
We have studied the phase transition of a real scalar $\varphi^4$ theory in four dimensions at lowest order in the 2PI formalism. Due to the existence of the Landau pole, this model needs to be considered in the presence of a cut-off lying below the scale of the
Landau pole. For values of the parameters such that the Landau pole is much larger than all other scales, renormalization ensure insensitivity of the results with respect to the cut-off scale already below the Landau scale. We have illustrated these questions at the level of the gap equation by investigating under which conditions the cut-off can be continued above the Landau pole to infinity without noticing any sensible change in the solution of the gap equation. We have
also studied the effective potential and proven analytically that if there is a phase transition, it is
weakly first order, confirming earlier studies based on a numerical evaluation or a high temperature expansion of the effective potential.\\
Let us end by emphasizing that even though we imposed the renormalization and consistency conditions in the symmetric phase, we could have imposed them in the broken phase as well. Because $\phi=0$ is not always accessible we need to impose these conditions at $\phi=v\neq 0$ (where $v$ is arbitrary for the moment). The bare mass $m_0^2$ and the bare coupling $\lambda_0$ are used to impose the renormalization conditions
\beq
\bar M^2_{\phi=v,\,T=0}=m^2>0 \quad {\rm and} \quad \bar V_{\phi=v,\,T=0}=\lambda>0\,.
\eeq
Unlike what happened at $\phi=0$ in the symmetric phase, the curvature $\hat M^2_{\phi=v,\,T=0}$ is not equal to $\bar M^2_{\phi=v,\,T=0}$. This truncation artifact can be overcome by using the bare coupling $\lambda_4$ to impose the consistency condition:
\beq\label{eq:consist2}
\hat M^2_{\phi=v,\,T=0}=\bar M^2_{\phi=v,\,T=0}\,.
\eeq
Notice then that the consistency condition:
\beq\label{eq:consist4}
\hat V_{\phi=v,\,T=0}=\bar V_{\phi=v,\,T=0}\,,
\eeq
fixes the value of $v$ in terms of $m^2$ and $\lambda$. An explicit calculation leads to:\footnote{This shows in particular that $\lambda<32\pi^2/3\approx 105$.}
\beq
x^2\equiv\frac{v^2}{m^2}=\frac{6}{\lambda-3\lambda^2/(32\pi^2)}\,,
\eeq
One can then check that $v$ cannot be interpreted as the value of $\bar\phi$ (minimum of the potential) at $T=0$ which, given the renormalization and consistency conditions, is such that $\bar\phi^2_{T=0}/m^2=3/\lambda$. Reversely, if one chooses $v$ as the value of $\bar\phi$ at $T=0$, which is the usual choice, then $v^2/m^2=3/\lambda$ and one cannot impose the consistency condition (\ref{eq:consist4}). In this later case, we have checked using similar arguments as above, that the transition cannot be second order.\\
The impossibility to impose simultaneously the consistency conditions (\ref{eq:consist2}) and (\ref{eq:consist4}) at the minimum of the potential needs to be regarded as a truncation artifact. It can be cured by exploiting the fact that in higher orders of approximation of the 2PI effective action, two bare masses $m_0^2$ and $m_2^2$ are usually needed to renormalize different quadratic divergences in $\hat M^2$ and $\bar M^2$, see for instance \cite{Arrizabalaga:2006hj}. It is true that in the Hartree approximation these quadratic divergences are equal, but we could still allow for different finite parts to $m^2_0$ and $m_2^2$ and use them to impose both consistency conditions at the minimum of the effective potential. Notice that this approach and the one involving only one bare mass are not completely equivalent. They both agree at leading order in $\lambda$ but differ beyond this order: in the first approach, some of the consistency conditions, which reflect an exact property of the theory are sacrificed, whereas in the second approach, what is sacrificed is the fact that in the exact theory there is only one bare mass. The second approach can be viewed as a renormalization improvement over the first one, for exact properties of the theory are reflected at the level of the renormalized quantities not at the level of the bare ones. It is then interesting to wonder how such an improvement affects the order of the transition. Interestingly, before exploiting the consistency condition (\ref{eq:consist4}) a second order phase transition is possible for parameters such that $v^2/m^2<1/\lambda$. However, the consistency condition (\ref{eq:consist4}) leads once again to a relation between $v^2/m^2$ and $\lambda$:
\beq
0=\frac{\lambda^4}{32\pi^2}\left(\frac{3\lambda}{32\pi^2}-1\right)\frac{v^6}{m^6}+\frac{3\lambda^3}{16\pi^2}\frac{v^4}{m^4}-\lambda \frac{v^2}{m^2}+3\,.
\eeq
For positive $\lambda$, this equation has a positive solution $v^2/m^2$ only for $\lambda<32\pi^2/3$. Moreover, it is possible to show that $v^2/m^2>3/\lambda$ and thus the region $v^2/m^2<1/\lambda$ cannot be accessed. For parameters such that $v^2/m^2>3/\lambda$, it is again possible to convince oneself that the transition cannot be of second order. Then, our conclusions on the nature of the phase transition in the Hartree approximation remain unchanged even within this extended renormalization scheme. The impossibility of a second order phase transition
is an artifact of the Hartree approximation. Therefore, it would be of
interest to see to what extend one could apply the analysis and
methods of the current work to more complicated approximations of the
2PI functional, in which there exist in the literature numerical
evidence for the occurrence of a second order phase transition \cite{Arrizabalaga:2006hj}.
Investigations in this direction are under way.
\acknowledgements{We would like to thank J.-P. Blaizot, G. Fej\H{o}s and J. Serreau for useful discussions on related topics.}
|
\section{QCD and Chiral Perturbation Theory}
\ \\
The Lagrangian of QCD is scale invariant, but its quantisation
singles out an intrinsic energy
$\Lambda_{\rm QCD}$, which sets the scale for the hadron spectrum.
Our daily life is dominated by low energy and therefore by the
lightest quark flavours,
{\it i.e.}\ quarks with masses $m_{\rm q} \ll \Lambda_{\rm QCD}$.
In the limit of vanishing quark masses, their left- and right-handed
spinor components ($\Psi_{L}$ and $\Psi_{R}$) decouple.
Thus the Lagrangian takes the structure
\begin{equation}
{\cal L}_{\rm QCD} = \bar \Psi_{L} D \Psi_{L} +\bar \Psi_{R} D \Psi_{R}
+ {\cal L}_{\rm gauge} \ ,
\end{equation}
where $D$ is the Dirac operator. For $N_{f}$ massless
quark flavours, this Lagrangian has the global symmetry
\be
U(N_{f})_{L} \otimes U(N_{f})_{R} =
\underbrace{SU(N_{f})_{L} \otimes SU(N_{f})_{R}}_{\rm
chiral~flavour~symmetry} \otimes
\underbrace{U(1)_{V}}_{\rm baryon~number~conservation}
\otimes \underbrace{U(1)_{A}}_{\rm axial~symmetry} \ .
\ee
Here we split off the phases; the (vectorial) symmetry under
simultaneous left- and right-handed phase rotation, $U(1)_{V}$,
corresponds to the conservation of the baryon number.
The remaining $U(1)_{A}$ symmetry --- for opposite
$L$ and $R$ phase rotations --- is the axial symmetry. That is
a symmetry of the classical theory, which breaks explicitly
under quantisation, {\it i.e.}\ it is anomalous. We are interested
in the remaining {\em chiral flavour symmetry,} which (in infinite
volume) breaks spontaneously,
\begin{equation}
SU(N_{f})_{L} \otimes SU(N_{f})_{R} \longrightarrow SU(N_{f})_{L+R} \ .
\end{equation}
{\em Chiral Perturbation Theory} ($\chi$PT) deals with an effective
Lagrangian in term of fields in the coset space of this spontaneous
symmetry breaking, $U(x) \in SU(N_{f})$ \cite{XPT}.
Thus it captures the lightest
degrees of freedom, which dominate low energy physics, in this case
given by $N_{f}^{2}-1$ Nambu-Goldstone bosons. The effective
chiral Lagrangian ${\cal L}_{\rm eff}$
embraces all terms which are compatible with the symmetries.
This concept also extends to the case where small quark masses are
added, $m_{\rm q} \gsim 0$. Then one deals with light pseudo
Nambu-Goldstone bosons, which are identified with the light mesons;
for $N_{f}=3$ this includes the pions, the kaons and the $\eta$-meson.
Here we consider the case $N_{f}=2$, so we only deal with the quark
flavours $u$ and $d$. We assume them to be degenerate, {\it i.e.}\
to have both the mass $m_{\rm q}$. In this case the field
$U(x) \in SU(2)$ describes the pion triplet.
The terms in ${\cal L}_{\rm eff}$
are ordered according to an energy hierarchy, which depends on the
number of derivatives and powers of $m_{\rm q}$.
Some of the first terms are
\bea
{\cal L}_{\rm eff} &=& \frac{1}{4} F^{2}_{\pi}\, {\rm Tr}
[ \partial_{\mu} U^{\dagger}\partial_{\mu} U ] + \frac{1}{2}
\Sigma m_{\rm q} \, {\rm Tr} [ U + U^{\dagger} ] \nn \\
&& - \frac{1}{4} l_{1} \left( {\rm Tr} [ \partial_{\mu} U^{\dagger}
\partial_{\mu} U ] \right)^{2} - \frac{1}{4} l_{2} \left( {\rm Tr}
[ \partial_{\mu} U^{\dagger} \partial_{\nu} U ] \right)^{2} \nn \\
&& - (l_{3} + l_{4}) \Big( \frac{\Sigma m_{\rm q}}{2 F_{\pi}} \Big)^{2}
\left( {\rm Tr} [ U + U^{\dagger} ] \right)^{2} +
l_{4} \frac{\Sigma m_{\rm q}}
{4 F_{\pi}} \, {\rm Tr} [\partial_{\mu} U^{\dagger} \partial_{\nu} U ]
\, {\rm Tr} [ U + U^{\dagger} ] + \dots
\label{Leff}
\eea
Each term comes with a coefficient, which is a free parameter
of the effective theory. These coefficients are known as
the {\em Low Energy Constants} (LECs). The leading LECs are the
pion decay constant $F_{\pi}$ (which was measured as
$F_{\pi} = 92.4 ~ {\rm MeV}$) and the chiral condensate $\Sigma$
(the order parameter for chiral symmetry breaking). The $l_{i}$
are sub-leading LECs. In the
chiral limit $m_{\rm q}=0$, only $F_{\pi}$ occurs in the leading order.
This determines an intrinsic scale $\Lambda_{\chi} = 4 \pi F_{\pi}
\simeq 1.2 ~ {\rm GeV}$, which adopts the r\^{o}le of $\Lambda_{\rm QCD}$
in $\chi$PT.
The LECs are of primary importance for low energy hadron physics.
To some extent they can be fixed from phenomenology. On the theoretical
side, they can only be determined from QCD as the underlying fundamental
theory. Since this refers to low energy, it is a non-perturbative
task, and therefore a challenge for lattice simulations. If one succeeds
in their determination, we arrive at a rather complete, QCD-based
formalism for low energy hadron physics.
\section{Pions in a finite volume: $p$-regime, $\epsilon$-regime and
$\delta$-regime}
\ \\
For a system of pions in a finite volume, say with
periodic boundary conditions and with a characteristic extent
$L$, the low energy expansion can be formulated in terms of the
dimensionless lowest non-zero momentum $p_{\mu}/\Lambda_{\chi}
\sim 1/(2 F_{\pi} L)$, and of the correlation length
$\xi = M_{\pi}^{-1}$ ($M_{\pi}$ is the pion mass).
Both $p_{\mu}$ and $M_{\pi}$
should be light compared to $\Lambda_{\chi}$. Depending on the
size and shape of the volume, one distinguishes various
regimes, with different counting rules for these expansion parameters:
\begin{itemize}
\item {\underline{$p$-Regime:}} This is the standard setting with
a large volume, $L \gg \xi$, and therefore small finite size
effects \cite{preg}.
In this case the counting is simply $O(1/L) = O(M_{\pi})$. From
the Lagrangian we can read off (on tree level) the
Gell-Mann--Oakes--Renner relation
\be \label{GMOR}
M_{\pi}^{2} = \frac{\Sigma}{F_{\pi}^{2}} \, m_{\rm q} \ .
\ee
\item {\underline{$\epsilon$-Regime:}} This regime refers to a small
box in Euclidean space \cite{epsreg1},
say $V = L^{4}$ with $L \llsim \xi$ (where
$\xi$ is still the would-be inverse pion mass in a large volume,
with the same quark mass; in Nature $\xi \simeq 1.5 ~ {\rm fm}$).
In this regime, the $\chi$PT counting rules read
$O(1/L) = O(m_{\rm q}) = O(M_{\pi}^{2} / \Lambda_{\chi})$,
unlike the $p$-regime counting.
Experimentally the $\epsilon$-regime is not accessible, but QCD
simulations in (or at least close to)
this regime are feasible. They are of interest
in particular because the LECs determined in this regime are
the same that occur in large volume. Therefore we can
extract physical information even from an unphysical regime.
Since this can be achieved with a modest lattice size, this
method is attractive from a practical perspective \cite{eps1}.
It has been intensively explored since 2003. It is difficult to
simulate safely inside this regime, but certain
properties, which are characteristic for the $\epsilon$-regime,
have been observed. Ref.\ \cite{eps3} provides a short overview.
\item {\underline{$\delta$-Regime:}} Here one deals with a
small {\em spatial} volume, but a large extent $T$ in Euclidean
time \cite{deltareg}, say
\be
L^{3} \times T \quad , \qquad L \llsim \xi \ll T \ .
\ee
This relation for $L$ and $T$ is depicted in Fig.\ \ref{deltaboxmap}
on the left (it is exactly opposite to the setting used in studies of
QCD at finite temperature). The counting rules for the
corresponding {\em $\delta$-expansion} are
\be
\frac{1}{\Lambda_{\chi} L} = O(\delta) \quad , \qquad
\frac{M_{\pi}}{\Lambda_{\chi}} \ , \ \frac{1}{\Lambda_{\chi} T}
= O(\delta^{3}) \ .
\ee
\end{itemize}
A map of these three regimes in terms of the
pion mass and the inverse extent in Euclidean time
(the temperature) is shown in Fig.\ \ref{deltaboxmap} on the right.
This article addresses the $\delta$-regime.
It is far less known and explored
than the $p$- and the $\epsilon$-regime, but it shares with the
latter the exciting property that physical LECs can be extracted
from an unphysical setting.
\begin{figure}[h!]
\begin{center}
\includegraphics[width=13pc]{deltavol.eps} \hspace*{1cm}
\includegraphics[width=17pc]{chiPTmap2.eps}
\end{center}
\caption{\label{deltaboxmap}On the left: an illustration of
a typical shape of a $\delta$-box, {\it i.e.}\ an anisotropic
finite volume where a pion gas can be treated by the
$\delta$-expansion. On the right: a schematic map of the applicability
domains of three different expansion rules of $\chi$PT, namely the
$p$- , the $\epsilon$- and the $\delta$-regime. The dashed lines
indicate regions where clearly one expansion holds; in the transition
zones between these regions various expansions could work more or less.}
\end{figure}
A further motivation for studying QCD in a ``$\delta$-box'' is that
its shape allows (approximately) for a simplified analytical treatment in
terms of 1d field theory, {\it i.e.}\ {\em quantum mechanics.}
In this case one considers a quantum rotator as described by the
1d $O(4)$ model, due to the local isomorphism between the orthogonal
group $O(4)$ and the chiral symmetry group $SU(2)_{L} \otimes
SU(2)_{R}$. Closely related systems have applications in solid state
physics, in particular regarding quantum anti-ferromagnets
\cite{solid,HasNie}.
Spontaneous symmetry breaking does not occur in a finite volume.
Therefore the pions\footnote{One might argue if the term ``pion''
is adequate in the $\delta$-regime. We find it acceptable and
convenient, but readers who disagree may simply denote $M_{\pi}^{R}$
(see below) as the ``mass gap''.}
--- {\it i.e.}\ the pseudo Nambu-Goldstone bosons ---
cannot become massless in the chiral limit $m_{\rm q}=0$, in contrast
to the infinite volume. We may consider a fixed box of a $\delta$-shape
and vary the quark mass: a large value of $m_{\rm q}$ implies a
large pion mass, so that we enter the $p$-regime and the
Gell-Mann--Oakes--Renner relation (\ref{GMOR}), $m_{q} \propto
M_{\pi}^{2}$, is approximated. For small
$m_{\rm q}$ the pion mass turns into a plateau, which ends in the
chiral limit at a {\em residual pion mass} $M_{\pi}^{R}$. This behaviour
is illustrated schematically in Fig.\ \ref{schema}.
\begin{figure}[h!]
\begin{center}
\includegraphics[width=13pc,angle=270]{Mpi.eps}
\end{center}
\caption{\label{schema} A qualitative picture of the expected
behaviour of the pion mass squared in a $\delta$-box.
For heavy quarks and pions we approximate
the $p$-regime relation $m_{q} \propto M_{\pi}^{2}$. For light
quarks the pion mass attains a plateau, and finally (in the chiral
limit $m_{\rm q}=0$) the residual value $M_{\pi}^{R}$.}
\end{figure}
The value of $M_{\pi}^{R}$ can be computed with
the $\delta$-expansion. The spectrum of the $O(4)$ quantum rotator
(a quantum mechanical particle on the sphere $S^{3}$) is given
by $E_{\ell} = \ell (\ell +2) / (2 \Theta)$, so the mass gap
amounts to $M_{\pi}^{R} = 3/ (2 \Theta)$. The challenge is now to
compute the moment of inertia $\Theta$. In his seminal
paper on the $\delta$-regime, H.\ Leutwyler gave
its value to leading order (LO) as $\Theta \approx F_{\pi}^{2} L^{3}$.
Thus the residual pion mass can be written as
\be \label{Mres}
M_{\pi}^{R} = \frac{3}{2 F_{\pi}^{2} L^{3} (1 + \Delta )} \ .
\ee
The shift $\Delta$ captures higher order corrections,
which are suppressed in powers of $1/ (F_{\pi}L)^{2}$.
They have been evaluated to next-to-leading order (NLO)
in Ref.\ \cite{HasNie}, and recently even to next-to-next-to-leading
order (NNLO) \cite{Has}, which yields
\be \label{Delta}
\Delta = \frac{0.4516 \dots}{F_{\pi}^{2} L^{2}} +
\frac{0.08843 \dots}{F_{\pi}^{4} L^{4}} \Big[ 1 - 0.1599 \dots
\Big( \ln (\Lambda_{1} L) + 4 \ln (\Lambda_{2} L) \Big) \Big] \ .
\ee
$\Lambda_{i}$ are scale parameters for the sub-leading LECs.
The latter are given at the scale of the physical pion mass as
\be \label{barli}
\bar l_{i} = \ln ( \Lambda_{i} / M_{\pi}^{\rm phys} )^{2} \ .
\ee
Even more recent papers addressed again the NNLO
of the $\delta$-expansion \cite{Weier}, and the corrections
due to finite $m_{\rm q}$ \cite{Wein}.
In the following we will discuss numerical results for $M_{\pi}^{R}$.
We see that a confrontation with the analytical prediction in eqs.\
(\ref{Mres}), (\ref{Delta}) could enable a new
determination of a set of LECs from first principles of QCD.
\section{Attempts to simulate QCD in the $\delta$-regime}
\ \\
The straight way to measure $M_{\pi}^{R}$ are simulations
directly {\em in} the $\delta$-regime. Since the $\delta$-box
differs from the lattice shapes in usual simulations, this requires
the special purpose generation of configurations.
Moreover, precise chirality is vital in this regime, hence
one is supposed to use a formulation of lattice quarks which
preserves chiral symmetry. Such lattice fermions are
known since the late 90ies,
but their simulation is extremely tedious,
in particular with dynamical quarks
({\it i.e.}\ keeping track of the fermion determinant in the
generation of gauge configurations).
We anticipate that so far there are no robust results
of simulations clearly inside the $\delta$-regime.
In this section we summarise the efforts that have been
carried out so far.
At the Symposium LATTICE2005 D.\ Hierl presented a first
attempt to simulate 2-flavour QCD in the $\delta$-regime \cite{HHHNS}.
That study used a truncated version of a chiral lattice Dirac
operator, so a first question is if the quality of approximate
chirality was sufficient for that purpose. Since this Dirac operator
is very complicated, that simulation was performed with a non-standard
algorithm, which probes the fermion determinant
with a stochastic estimator. The spatial volume
was $\approx (1.2 ~ {\rm fm})^{3}$, and the results for
$M_{\pi}$ at small quark masses agreed well with the LO
of the $\delta$-expansion, {\it i.e.}\ eq.\ (\ref{Mres})
at $\Delta =0$ (and with the phenomenological value of $F_{\pi}$).
In 2007 the QCDSF Collaboration generated a new set of data, which
have not been published. They were obtained with dynamical
overlap quarks, which are exactly chiral. This simulation used the
Hybrid Monte Carlo algorithm ({\it i.e.}\ the reliable standard
algorithm). The lattice had a modest size of $8^{3} \times 16$ sites,
and the spatial box length was again $L \approx 1.2 ~ {\rm fm}$.
At first sight the results seemed to look
fine: for decreasing $m_{\rm q}$ we saw a transition from a
Gell-Mann--Oakes--Renner type behaviour towards a plateau. Its value
agreed with the chirally extrapolated value of Ref.\ \cite{HHHNS},
and therefore also with the LO of eq.\ (\ref{Mres}).
Unfortunately, this is {\em not} the end of the story. If we proceed
to the NLO correction, {\it i.e.}\ if we include the first term
of $\Delta$ given in eq.\ (\ref{Delta}), the predicted value for
$M_{\pi}^{R}$ decreases drastically in this small volume --- from
$782.5 ~{\rm MeV}$ down to $321.7 ~{\rm MeV}$ --- and the agreement
with the above data is gone. Considering this dramatic effect of
the NLO correction one might worry that the $\delta$-expansion
could converge only very slowly in this small box,
and such simulations are not instructive at all. However,
adding also the NNLO correction
alters the NLO result only a little --- to
$(336.3 \pm 7.6) ~{\rm MeV}$ --- so it is reasonable to assume
the $\delta$-expansion to be already well converged.\footnote{To
obtain this theoretically predicted value, we inserted the LECs as
far as they are known. In particular the sub-leading LECs
$\bar l_{1} = -0.4 \pm 0.6$ and $\bar l_{2} = 4.1 \pm 0.1$
are taken from Ref.\ \cite{CGL}, along with their
uncertainties, which imply the uncertainty in the NNLO value of
$M_{\pi}^{R}$.\label{lifoot}}
Thus simulations in this small box {\em can} be useful
in view of a confrontation with the analytical predictions for
$M_{\pi}^{R}$. In fact a second sequence of runs by the QCDSF
Collaboration (performed in 2008) yielded data much closer to
the NNLO prediction. They involved 5 quark masses; the
results for the physical lattice spacing $a$ and the pion mass
are given in Table \ref{run2dat}.
Since $a$ varies for the different values of $m_{\rm q}$,
the box length $L=8a$ and the prediction for $M_{\pi}^{R}$ vary
as well.
In Fig.\ \ref{deltadat}
we compare the numerically measured pion masses and the corresponding
NNLO $\delta$-expansion results. They are compatible, in contrast to
the earlier data, which were probably not well thermalised.
In these new runs thermalisation is accomplished, but the ratio
$T/L=2$ is still modest. We therefore proceeded to a
$8^{3}\times 32$ lattice, where our runs are ongoing.
Still, the quality of agreement with the $\delta$-expansion
that we observed already on the $8^{3}\times 16$ lattice is impressive.
\begin{table}
\begin{center}
\begin{tabular}{|c||c|c|c|}
\hline
$m_{\rm q}$ & $a$ [fm] & $M_{\pi}$ & $M_{\pi}^{R}$ (NNLO) \\
\hline
0.008 & 0.102(9)& $(535.8 \pm 54.5)$ MeV & $(510.3 \pm 30.7)$ MeV \\
\hline
0.01 & 0.104(8) & $(580.5 \pm 47.2)$ MeV & $(505.8 \pm 30.3)$ MeV \\
\hline
0.03 & 0.114(8) & $(444.8 \pm 31.4)$ MeV & $(476.2 \pm 33.4)$ MeV \\
\hline
0.04 & 0.111(8) & $(476.4 \pm 34.7)$ MeV & $(486.2 \pm 32.6)$ MeV \\
\hline
0.05 & 0.111(9) & $(511.9 \pm 43.9)$ MeV & $(486.2 \pm 35.0)$ MeV \\
\hline
\end{tabular}
\end{center}
\caption{The results by QCDSF Collaboration of the year 2008, on
a $8^{3} \times 16$ lattice with 5 values of the mass $m_{\rm q}$
for the two degenerate dynamical overlap quark flavours.
We display the measured values of the physical lattice spacing $a$
and the pion mass, as well as the
NNLO prediction for the residual pion mass $M_{\pi}^{R}$.
(The errors capture the uncertainty in $a$, in the pion mass in
lattice units, and in the LECs $\bar l_{1}$ and $\bar l_{2}$,
cf.\ eq.\ (\ref{l1234}).)}
\label{run2dat}
\vspace{-5mm}
\end{table}
\begin{figure}[h!]
\begin{center}
\vspace*{-6mm}
\includegraphics[width=17.9pc,angle=270]{pionmassesdelta.eps}
\end{center}
\vspace*{-4mm}
\caption{An illustration of our results given in Table \ref{run2dat}:
we show the numerically measured pion mass
and the theoretical residual mass $M_{\pi}^{R}$, as predicted by the
NNLO $\delta$-expansion. (The latter actually refer to the chiral
limit, but its value depends on $L$ and therefore
on the lattice spacing $a$, which varies for the simulations
at different $m_{\rm q}$.) We observe good agreement between the
measured and the predicted values.}
\label{deltadat}
\vspace*{-5mm}
\end{figure}
\section{Residual pion mass by an extrapolation from the
$p$-regime}
\ \\
In this section we proceed to a different approach.
It is based on simulation results in the $p$-regime (up to the
transition zone),
which are then extrapolated
towards the $\delta$-regime. Details of this study are given in
Ref.\ \cite{QCDSFdelta}. Also in this framework we consider it essential
to use dynamical quarks, but we do not insist on exact chiral symmetry
in the $p$-regime. Hence we used Wilson
fermions, which is an established standard lattice fermion formulation,
in a form which corrects $O(a)$ lattice artifacts. Thus the simulation
was much faster, and we could tackle much larger lattices than those
mentioned in the Section 3; our data reported below were
obtained on three lattice sizes:
$ 24^{3} \times 48 \ , \ 32^{3} \times 64 $ and $40^{3} \times 64 \ .$
On the other hand, this lattice regularisation breaks
the chiral symmetry explicitly, so that additive mass
renormalisation sets in. Nevertheless we were able
to attain very light pion masses.
For the gauge part we used the standard plaquette lattice action.
Our simulations were carried out at two values for the strong
gauge coupling $g_{\rm s}$, respectively the parameter
$\beta = 6 / g_{\rm s}^{2}$. We determined the physical
lattice spacing $a$ from the measured nucleon mass,
which revealed that we were dealing with fine lattices,
\be
\beta = 5.29 \ \to \ a \simeq 0.075 ~ {\rm fm} \quad ,
\quad
\beta = 5.4 \ \to \ a \simeq 0.067 ~ {\rm fm} \quad .
\ee
Thus the spatial size was in the range
$L \simeq 1.6 ~{\rm fm} \dots 3.0 ~{\rm fm} .$
Due to the additive mass renormalisation, we could not refer to the
bare quark mass anymore. We measured the current
quark mass by means of the PCAC relation,
\be
m_{\rm q} = \frac{\langle
\partial_{4} A_{4} (\vec 0 , x_{4}) P(0) \rangle}
{\langle P (\vec 0 , x_{4}) P(0) \rangle} \ ,
\ee
where $P$ is the pseudoscalar density, and $A_{\mu}$ is the
axial current.
We observed practically no finite size effects on $a$
and on $m_{\rm q}$, but we did see a striking $L$-dependence of
$M_{\pi}$, as expected. These quantities were found in the range
\be
m_{\rm q} = 3.60 ~ {\rm MeV} \dots 231 ~ {\rm MeV} \quad , \quad
M_{\pi} = 174 ~ {\rm MeV} \dots 1.52 ~ {\rm GeV} \quad
, \quad M_{\pi} L = 2.7 \dots 9.7 \ .
\ee
The latter confirms that our data range from the deep $p$-regime
to the transition zone. A few missing data points
have been completed with a (lengthy) formula for exponentially
suppressed finite size effects \cite{ML,CDH}
within the $p$-regime, which holds up to $O(p^{4})$.
This formula involves the renormalised sub-leading LECs
$\bar l_{i}, \ i=1 \dots 4$, see eqs.\ (\ref{Leff}) and (\ref{barli}).
Well established phenomenological values
were composed in Ref.\ \cite{CGL}
($\bar l_{1}, \ \bar l_{2}$ were anticipated in footnote \ref{lifoot}),
\be \label{l1234}
\bar l_{1} = -0.4 \pm 0.6 \quad , \quad \bar l_{2} = 4.1 \pm 0.1
\quad , \quad \bar l_{3} = 2.9 \pm 2.4
\quad , \quad \bar l_{4} = 4.4 \pm 0.4 \quad .
\ee
They were estimated from
$\pi \pi$ scattering data, and in particular the
$\bar l_{4}$ value is based on the scalar pion form factor.
The dark horse in this context is $\bar l_{3}$:
we replaced the above value by $\bar l_{3} \approx 4.2$,
which we obtained in our study, see below.
Our measured and interpolated data are given in
Ref.\ \cite{QCDSFdelta}. We extrapolated them towards
the $\delta$-regime and extracted in particular a value for
$M_{\pi}^{R}$ based on the relatively simple
chiral extrapolation formula
\be \label{fitfun}
M_{\pi} (L)^{2} = M_{\pi}^{R~2} + C_{1} m_{\rm q}
\left[ 1 + C_{2} m_{\rm q} \ln (C_{3} m_{\rm q}) \right] \ ,
\ee
which interpolates between the $O(p^{4})$ correction formula
(for large $L M_{\pi}$) and the chiral limit \cite{CGL}.
The $C_{i}$ and $M_{\pi}^{R}$ are treated as free parameters
to be fixed by the fit. Our data and the fits for
$\beta = 5.29$ and for $\beta = 5.4$ are shown in Figs.\
\ref{b529} and \ref{b540}, respectively. These plots also
illustrate the extrapolation result for $M_{\pi}^{R}$
in the chiral limit.
\begin{figure}[h!]
\begin{center}
\includegraphics[width=17.3pc,angle=0]{mpimq_529_min2.ps}
\includegraphics[width=17.3pc,angle=0]{mpimq_529_40_min2x.ps}
\end{center}
\vspace*{-2mm}
\caption{Our chiral extrapolation referring to
the lattice spacing $a \simeq 0.075 ~ {\rm fm}$ (corresponding
to $\beta =5.29$) with $L \simeq 1.8 ~ {\rm fm}$ (on the left)
and $L \simeq 3.0 ~ {\rm fm}$ (on the right).
The data points show the measured pion mass
squared against the quark mass in the $p$-regime (the Sommer
scale parameter $r_{0}= 0.467 ~ {\rm fm}$ is employed to
convert them into dimensionless units). The curve is the fit
according to eq.\ (\ref{fitfun}). It ends in the chiral limit,
where we illustrate the extrapolated value for $M_{\pi}^{R}$ and its
error.}
\label{b529}
\end{figure}
\begin{figure}[h!]
\begin{center}
\vspace*{-1.3cm}
\includegraphics[width=17.3pc,angle=0]{mpimq_540_24_min2x.ps}
\includegraphics[width=17.3pc,angle=0]{mpimq_540_min2x.ps}
\end{center}
\vspace*{-2mm}
\caption{The same as Fig.\ \ref{b529}, but now for the finer
lattices with $a \simeq 0.067 ~ {\rm fm}$ (corresponding
to $\beta =5.4$), such that $L \simeq 1.6 ~ {\rm fm}$ (on the left)
and $L \simeq 2.1 ~ {\rm fm}$ (on the right).}
\vspace*{-2mm}
\label{b540}
\end{figure}
Our main result is shown in Fig.\ \ref{finale}. It compares
the extrapolated values for $M_{\pi}^{R}$ with the predictions
based on the $\delta$-expansion to LO, NLO and NNLO, as a function
of $L$. The latter two predictions are very close to each other
for the volumes under consideration, so we can again assume the
expansion to be well converged.
\begin{figure}[h!]
\begin{center}
\includegraphics[width=19pc,angle=270]{m_pi_res2.eps}
\end{center}
\caption{A comparison of our extrapolated values for the
residual pion mass $M_{\pi}^{R}$ with the predictions
by the $\delta$-expansion (LO, NLO, NNLO), as functions of
the spatial box size $L$. We observe a remarkably good agreement.
In our largest volume, the extrapolated $M_{\pi}^{R}$ is located
below the physical pion mass, and in the region $M_{\pi}^{R} L < 1$,
which can be viewed as the domain of the $\delta$-regime.}
\label{finale}
\end{figure}
The extrapolation results for $M_{\pi}^{R}$
reach down to values even below the physical pion mass.
The plot also shows the line where the product $M_{\pi}L$
decreases below 1; this can be roughly considered as
the boundary of the $\delta$-regime. Our extrapolations
lead close to this boundary, and in one case {\em into} the
$\delta$-regime.\footnote{In that case, the extrapolation
should actually turn into a plateau at tiny $m_{\rm q}$,
but this would hardly change the result.}
In particular this plot shows that the extrapolated masses
$M_{\pi}^{R}$ match this curve
remarkably well. It would not have been obvious to predict this
feature, because our data were obtained in a regime where the
basis of the $\delta$-expansion (frozen spatial degrees of freedom)
does not hold.
This suggests that the extrapolation formula (\ref{fitfun}) applies
well in a sizable domain.
\section{Evaluation of Low Energy Constants}
\ \\
By fitting our pion mass data according to eq.\ (\ref{fitfun}),
we obtained results for the four free fitting parameters.
In Section 4 we discussed the results for $M_{\pi}^{R}$ that
we obtained in this way. Moreover, the results for $C_{1}$,
$C_{2}$ and $C_{3}$ can be used to evaluate the notorious
sub-leading LEC $\bar l_{3}$ \cite{QCDSFdelta}. This is how
we obtained our value that we anticipated in Section 4,
\be
\bar l_{3} = 32 \pi^{2} F_{\pi}^{2} \, \frac{C_{2}}{C_{1}} \,
\ln \frac{C_{1}}{C_{3} M_{\pi}^{{\rm phys}~2}}
= 4.2 \pm 0.2 \ .
\ee
(The error given here emerges from the fits, the additional systematic
error would be hard to estimate).
In view of other results in the literature, this value is in
the upper region. An overview has been presented in Ref.\
\cite{Latdat}, in particular in Table 11 and Figure 9.
That overview estimates the world average as $\bar l_{3} = 3.3 (7)$.
In the previous considerations of Sections 4 and 5, we
inserted the phenomenological value of the pion decay constant,
$F_{\pi} = 92.4 ~{\rm MeV}$. Alternatively, we could also
treat $F_{\pi}$ as a free parameter to be determined by the fits.
In particular, matching $M_{\pi}^{R}$ this yields
in the chiral limit \cite{QCDSFdelta}
\be
\left. F_{\pi}^{\rm numerical} \right\vert_{m_{\rm q}=0} = 78^{+14}_{-10}
~{\rm MeV} \ ,
\ee
which seems a bit low. However, effective field theory considerations
suggest that the value of $F_{\pi}$ in the chiral limit
should indeed be below the physical value; Ref.\ \cite{Latdat}
estimates $F_{\pi} \vert_{m_{\rm q}=0} \approx 86 ~ {\rm MeV}$.
\section{Conclusions}
\ \\
The $\delta$-regime refers to a system of pions in a finite
box, typically of the shape $L^{3} \times T$ with
$L \llsim M_{\pi}^{-1} \ll T$. It can be treated by Chiral
Perturbation Theory with suitable counting rules,
the $\delta$-expansion. In particular this predicts
the residual pion mass in the chiral limit, $M_{\pi}^{R}$,
which has been computed recently to NNLO \cite{Has}.
First attempts to simulate 2-flavour QCD in the $\delta$-regime
were confronted with technical difficulties, in particular
thermalisation problems. Nevertheless we obtained good results for the
pion mass with light quarks on a small lattice of size
$8^{3} \times 16$, with $L \simeq (0.82 \dots 0.91) ~{\rm fm}$.
In another pilot study we measured pion masses in the $p$-regime
(up to the transition region) and extrapolated them towards the
$\delta$-regime. This yields numerical results for $M_{\pi}^{R}$,
which agree remarkably well with the predictions by the
$\delta$-expansion. That expansion is based on assumptions,
which do not hold in the regime where the data were obtained.
Therefore it would not have been obvious to predict the observed
agreement. This comparison can be viewed as a numerical
experiment, which led to an interesting observation.
Our pion mass fits from the $p$- towards the $\delta$-regime
also fix some further constants, which allow
for a new determination of the (mysterious) sub-leading
Low Energy Constant $\bar l_{3}$; we obtained $\bar l_{3} = 4.2(2)$.
This is somewhat above the average of the phenomenological
and numerical estimates in the literature. The constants
$\bar l_{1}, \dots , \bar l_{4}$ are relevant in the effective
description of $\pi \pi$ scattering.
Finally we also considered the option to treat $F_{\pi}$
as a free parameter. In the chiral limit the fits lead to a
value somewhat below the phenomenological $F_{\pi}$.
Robust numerical results, measured manifestly inside the $\delta$-regime,
are still outstanding.\\
\vspace*{-1mm}
\noindent
{\bf Acknowledgements :} W.B.\ thanks for kind hospitality
at the University of Bern, where this talk was written up.
The simulations for this project were performed on the SGI Altix 4700
at LRZ (Munich), the IBM BlueGeneL and BlueGeneP at NIC (J\"{u}lich),
the BlueGeneL at EPCC (Edinburgh), and the apeNEXT at NIC/DESY
(Zeuthen). We also thank for financial support by the EU Integrated
Infrastructure Initiative ``HadronPhysics2'' and by the DFG under
contract SFB/TR 55 (Hadron Physics from Lattice QCD). J.M.Z.\ is supported
through the UKs STFC Advanced Fellowship Program under contract
ST/F009658/1.
\vspace*{-2mm}
\section*{References}
|
\section{Introduction}
The advent of ubiquitous large-scale distributed systems advocates that tolerance to various kinds of faults and hazards must be included from the very early design of such systems. \emph{Self-stabilization}~\cite{D74j,D00b,T09bc} is a versatile technique that permits forward recovery from any kind of \emph{transient} faults, while \emph{Byzantine Fault-tolerance}~\cite{LSP82j} is traditionally used to mask the effect of a limited number of \emph{malicious} faults. Making distributed systems tolerant to both transient and malicious faults is appealing yet proved difficult~\cite{DW04j,DD05c,NA02c} as impossibility results are expected in many cases.
\paragraph{Related Works}
A promizing path towards multitolerance to both transient and Byzantine faults is \emph{Byzantine containment}. For \emph{local} tasks (\emph{i.e.} tasks whose correctness can be checked locally, such as vertex coloring, link coloring, or dining philosophers), the notion of \emph{strict stabilization} was proposed~\cite{NA02c,MT07j}. Strict stabilization guarantees that there exists a \emph{containment radius} outside which the effect of permanent faults is masked, provided that the problem specification makes it possible to break the causality chain that is caused by the faults. As many problems are not local, it turns out that it is impossible to provide strict stabilization for those. To circumvent impossibility results, the weaker notion of \emph{strong stabilization} was proposed~\cite{MT06cb,DMT11j}: here, correct nodes outside the containment radius may be perturbated by the actions of Byzantine node, but only a finite number of times.
Recently, the idea of generalizing strict and strong stabilization to an area that depends on the graph topology and the problem to be solved rather than an arbitrary fixed containment radius was proposed~\cite{DMT10ca,DMT10cd} and denoted by \emph{topology aware} strict (and strong) stabilization. When maximizable metric trees are considered, \cite{DMT10ca} proposed an optimal (with respect to impossibility results) protocol for topology-aware strict stabilization, and for the simpler case of breath-first-search metric trees, \cite{DMT10cd} presented a protocol that is optimal both with respect to strict and strong variants of topology-aware stabilization. The case of optimality for topology-aware strong stabilization in the general maximal metric case remains open.
\paragraph{Our Contribution}
In this paper, we investigate the possibility of topology-aware strong stabilization for tasks that are global (\emph{i.e.} for with there exists a causality chain of size $r$, where $r$ depends on $n$ the size of the network), and focus on the maximum metric tree problem. In more details, we provide two necessary conditions to perform Byzantine containment for maximum metric tree construction. First, we characterize a specific class of maximizable metrics (which includes breath-first-search and shortest path metrics) that prevents the exitence of strong stabilizing solutions. Then, we generalize an impossibilty result of \cite{DMT10cd} that provides a lower bound on the containmemt area for topology-aware strong stabilization.
\section{Model and Definitions}
\subsection{State Model}
A \emph{distributed system} $S=(P,L)$ consists of a set $P=\{v_1,v_2,\ldots,v_n\}$ of processes and a set $L$ of bidirectional communication links (simply called links). A link is an unordered pair of distinct processes. A distributed system $S$ can be regarded as a graph whose vertex set is $P$ and whose link set is $L$, so we use graph terminology to describe a distributed system $S$. We use the following notations: $n=|P|$, $m=|L|$ and $d(u,v)$ denotes the shortest path between two processes $u$ and $v$ (\emph{i.e} the length of the shortest path between $u$ and $v$).
Processes $u$ and $v$ are called \emph{neighbors} if $(u,v)\in L$. The set of neighbors of a process $v$ is denoted by $N_v$. We do not assume existence of a unique identifier for each process. Instead we assume each process can distinguish its neighbors from each other by locally labelling them.
In this paper, we consider distributed systems of arbitrary topology. We assume that a single process is distinguished as a \emph{root}, and all the other processes are identical. We adopt the \emph{shared state model} as a communication model in this paper, where each process can directly read the states of its neighbors.
The variables that are maintained by processes denote process states. A process may take actions during the execution of the system. An action is simply a function that is executed in an atomic manner by the process. The action executed by each process is described by a finite set of guarded actions of the form $\langle$guard$\rangle\longrightarrow\langle$statement$\rangle$. Each guard of process $u$ is a boolean expression involving the variables of $u$ and its neighbors.
A global state of a distributed system is called a \emph{configuration} and is specified by a product of states of all processes. We define $C$ to be the set of all possible configurations of a distributed system $S$. For a process set $R \subseteq P$ and two configurations $\rho$ and $\rho'$, we denote $\rho \stackrel{R}{\mapsto} \rho'$ when $\rho$ changes to $\rho'$ by executing an action of each process in $R$ simultaneously. Notice that $\rho$ and $\rho'$ can be different only in the states of processes in $R$. For completeness of execution semantics, we should clarify the configuration resulting from simultaneous actions of neighboring processes. The action of a process depends only on its state at $\rho$ and the states of its neighbors at $\rho$, and the result of the action reflects on the state of the process at $\rho '$.
We say that a process is \emph{enabled} in a configuration $\rho$ if the guard of at least one of its actions is evaluated as true in $\rho$.
A \emph{schedule} of a distributed system is an infinite sequence of process sets. Let $Q=R^1, R^2, \ldots$ be a schedule, where $R^i \subseteq P$ holds for each $i\ (i \ge 1)$. An infinite sequence of configurations $e=\rho_0,\rho_1,\ldots$ is called an \emph{execution} from an initial configuration $\rho_0$ by a schedule $Q$, if $e$ satisfies $\rho_{i-1} \stackrel{R^i}{\mapsto} \rho_i$ for each $i\ (i \ge 1)$. Process actions are executed atomically, and we distinguish some properties on the scheduler (or daemon). A \emph{distributed daemon} schedules the actions of processes such that any subset of processes can simultaneously execute their actions. We say that the daemon is \emph{central} if it schedules action of only one process at any step. The set of all possible executions from $\rho_0\in C$ is denoted by $E_{\rho_0}$. The set of all possible executions is denoted by $E$, that is, $E=\bigcup_{\rho\in C}E_{\rho}$. We consider \emph{asynchronous} distributed systems where we can make no assumption on schedules.
In this paper, we consider (permanent) \emph{Byzantine faults}: a Byzantine process (\emph{i.e.} a Byzantine-faulty process) can make arbitrary behavior independently from its actions. If $v$ is a Byzantine process, $v$ can repeatedly change its variables arbitrarily. For a given execution, the number of faulty processes is arbitrary but we assume that the root process is never faulty.
\subsection{Self-Stabilizing Protocols Resilient to Byzantine Faults}
Problems considered in this paper are so-called \emph{static problems}, \emph{i.e.} they require the system to find static solutions. For example, the spanning-tree construction problem is a static problem, while the mutual exclusion problem is not. Some static problems can be defined by a \emph{specification predicate} (shortly, specification), $spec(v)$, for each process $v$: a configuration is a desired one (with a solution) if every process satisfies $spec(v)$. A specification $spec(v)$ is a boolean expression on variables of $P_v~(\subseteq P)$ where $P_v$ is the set of processes whose variables appear in $spec(v)$. The variables appearing in the specification are called \emph{output variables} (shortly, \emph{O-variables}). In what follows, we consider a static problem defined by specification $spec(v)$.
A \emph{self-stabilizing protocol} (\cite{D74j}) is a protocol that eventually reaches a \emph{legitimate configuration}, where $spec(v)$ holds at every process $v$, regardless of the initial configuration. Once it reaches a legitimate configuration, every process never changes its O-variables and always satisfies $spec(v)$. From this definition, a self-stabilizing protocol is expected to tolerate any number and any type of transient faults since it can eventually recover from any configuration affected by the transient faults. However, the recovery from any configuration is guaranteed only when every process correctly executes its action from the configuration, \emph{i.e.}, we do not consider existence of permanently faulty processes.
When (permanent) Byzantine processes exist, Byzantine processes may not satisfy $spec(v)$. In addition, correct processes near the Byzantine processes can be influenced and may be unable to satisfy $spec(v)$. Nesterenko and Arora~\cite{NA02c} define a \emph{strictly stabilizing protocol} as a self-stabilizing protocol resilient to unbounded number of Byzantine processes.
Given an integer $c$, a \emph{$c$-correct process} is a process defined as follows.
\begin{definition}[$c$-correct process]
A process is $c$-correct if it is correct (\emph{i.e.} not Byzantine) and located at distance more than $c$ from any Byzantine process.
\end{definition}
\begin{definition}[$(c,f)$-containment]
\label{def:cfcontained}
A configuration $\rho$ is \emph{$(c,f)$-contained} for specification $spec$ if, given at most $f$ Byzantine processes, in any execution starting from $\rho$, every $c$-correct process $v$ always satisfies $spec(v)$ and never changes its O-variables.
\end{definition}
The parameter $c$ of Definition~\ref{def:cfcontained} refers to the \emph{containment radius} defined in \cite{NA02c}. The parameter $f$ refers explicitly to the number of Byzantine processes, while \cite{NA02c} dealt with unbounded number of Byzantine faults (that is $f\in\{0\ldots n\}$).
\begin{definition}[$(c,f)$-strict stabilization]
\label{def:cfstabilizing}
A protocol is \emph{$(c,f)$-strictly stabilizing} for specification $spec$ if, given at most $f$ Byzantine processes, any execution $e=\rho_0,\rho_1,\ldots$ contains a configuration $\rho_i$ that is $(c,f)$-contained for $spec$.
\end{definition}
An important limitation of the model of \cite{NA02c} is the notion of $r$-\emph{restrictive} specifications. Intuitively, a specification is $r$-restrictive if it prevents combinations of states that belong to two processes $u$ and $v$ that are at least $r$ hops away. An important consequence related to Byzantine tolerance is that the containment radius of protocols solving those specifications is at least $r$. For some (global) problems $r$ can not be bounded by a constant. In consequence, we can show that there exists no $(c,1)$-strictly stabilizing
protocol for such a problem for any (finite) integer $c$.
\paragraph{Strong stabilization} To circumvent such impossibility results, \cite{DMT11j} defines a weaker notion than the strict stabilization. Here, the requirement to the containment radius is relaxed, \emph{i.e.} there may exist processes outside the containment radius that invalidate the specification predicate, due to Byzantine actions. However, the impact of Byzantine triggered action is limited in times: the set of Byzantine processes may only impact processes outside the containment radius a bounded number of times, even if Byzantine processes execute an infinite number of actions.
In the following of this section, we recall the formal definition of strong stabilization adopted in \cite{DMT11j}. From the states of $c$-correct processes, \emph{$c$-legitimate configurations} and \emph{$c$-stable configurations} are defined as follows.
\begin{definition}[$c$-legitimate configuration]
A configuration $\rho$ is $c$-legitimate for \emph{spec} if every $c$-correct process $v$ satisfies $spec(v)$.
\end{definition}
\begin{definition}[$c$-stable configuration]
A configuration $\rho$ is $c$-stable if every $c$-correct process never changes the values of its O-variables as long as Byzantine processes make no action.
\end{definition}
Roughly speaking, the aim of self-stabilization is to guarantee that a distributed system eventually reaches a $c$-legitimate and $c$-stable configuration. However, a self-stabilizing system can be disturbed by Byzantine processes after reaching a $c$-legitimate and $c$-stable configuration. The \emph{$c$-disruption} represents the period where $c$-correct processes are disturbed by Byzantine processes and is defined as follows
\begin{definition}[$c$-disruption]
A portion of execution $e=\rho_0,\rho_1,\ldots,\rho_t$ ($t>1$) is a $c$-disruption if and only if the following holds:
\begin{enumerate}
\item $e$ is finite,
\item $e$ contains at least one action of a $c$-correct process for changing the value of an O-variable,
\item $\rho_0$ is $c$-legitimate for \emph{spec} and $c$-stable, and
\item $\rho_t$ is the first configuration after $\rho_0$ such that $\rho_t$ is $c$-legitimate for \emph{spec} and $c$-stable.
\end{enumerate}
\end{definition}
Now we can define a self-stabilizing protocol such that Byzantine processes may only impact processes outside the containment radius a bounded number of times, even if Byzantine processes execute an infinite number of actions.
\begin{definition}[$(t,k,c,f)$-time contained configuration]
A configuration $\rho_0$ is $(t,k,c,f)$-time contained for \emph{spec} if given at most $f$ Byzantine processes, the following properties are satisfied:
\begin{enumerate}
\item $\rho_0$ is $c$-legitimate for \emph{spec} and $c$-stable,
\item every execution starting from $\rho_0$ contains a $c$-legitimate configuration for \emph{spec} after which the values of all the O-variables of $c$-correct processes remain unchanged (even when Byzantine processes make actions repeatedly and forever),
\item every execution starting from $\rho_0$ contains at most $t$ $c$-disruptions, and
\item every execution starting from $\rho_0$ contains at most $k$ actions of changing the values of O-variables for each $c$-correct process.
\end{enumerate}
\end{definition}
\begin{definition}[$(t,c,f)$-strongly stabilizing protocol]
A protocol $A$ is $(t,c,f)$-strongly stabilizing if and only if starting from any arbitrary configuration, every execution involving at most $f$ Byzantine processes contains a $(t,k,c,f)$-time contained configuration that is reached after at most $l$ rounds. Parameters $l$ and $k$ are respectively the $(t,c,f)$-stabilization time and the $(t,c,f)$-process-disruption times of $A$.
\end{definition}
Note that a $(t,k,c,f)$-time contained configuration is a $(c,f)$-contained configuration when $t=k=0$, and thus, $(t,k,c,f)$-time contained configuration is a generalization (relaxation) of a $(c,f)$-contained configuration. Thus, a strongly stabilizing protocol is weaker than a strictly stabilizing one (as processes outside the containment radius may take incorrect actions due to Byzantine influence). However, a strongly stabilizing protocol is stronger than a classical self-stabilizing one (that may never meet their specification in the presence of Byzantine processes).
The parameters $t$, $k$ and $c$ are introduced to quantify the strength of fault containment, we do not require each process to know the values of the parameters.
\paragraph{Topology-aware Byzantine resilience} We saw previously that there exist a number of impossibility results on strict stabilization due to the notion of $r$-restrictives specifications. To circumvent this impossibility result, we describe here another weaker notion than the strict stabilization: the \emph{topology-aware strict stabilization} (denoted by TA strict stabilization for short) introduced by \cite{DMT10ca}. Here, the requirement to the containment radius is relaxed, \emph{i.e.} the set of processes which may be disturbed by Byzantine ones is not reduced to the union of $c$-neighborhood of Byzantine processes (\emph{i.e.} the set of processes at distance at most $c$ from a Byzantine process) but can be defined depending on the graph topology and Byzantine processes location.
In the following, we give formal definition of this new kind of Byzantine containment. From now, $B$ denotes the set of Byzantine processes and $S_B$ (which is function of $B$) denotes a subset of $V$ (intuitively, this set gathers all processes which may be disturbed by Byzantine processes).
\begin{definition}[$S_{B}$-correct node]
A node is \emph{$S_{B}$-correct} if it is a correct node (\emph{i.e.} not Byzantine) which not belongs to $S_{B}$.
\end{definition}
\begin{definition}[$S_{B}$-legitimate configuration]
A configuration $\rho$ is \emph{$S_{B}$-legitimate} for $spec$ if every $S_{B}$-correct node $v$ is legitimate for $spec$ (\emph{i.e.} if $spec(v)$ holds).
\end{definition}
\begin{definition}[$(S_{B},f)$-topology-aware containment]
\label{def:SfTAcontained}
A configuration $\rho_{0}$ is \emph{$(S_{B},f)$-topology-aware contained} for specification $spec$ if, given at most $f$ Byzantine processes, in any execution $e=\rho_0,\rho_1,\ldots$, every configuration is $S_{B}$-legitimate and every $S_B$-correct process never changes its O-variables.
\end{definition}
The parameter $S_{B}$ of Definition~\ref{def:SfTAcontained} refers to the \emph{containment area}. Any process which belongs to this set may be infinitely disturbed by Byzantine processes. The parameter $f$ refers explicitly to the number of Byzantine processes.
\begin{definition}[$(S_{B},f)$-topology-aware strict stabilization]
\label{def:SfTAStrictstabilizing}
A protocol is \emph{$(S_{B},f)$-topology-aware strictly stabilizing} for specification $spec$ if, given at most $f$ Byzantine processes, any execution $e=\rho_0,\rho_1,\ldots$ contains a configuration $\rho_i$ that is $(S_{B},f)$-topology-aware contained for $spec$.
\end{definition}
Note that, if $B$ denotes the set of Byzantine processes and $S_{B}=\left\{v\in V|\underset{b\in B}{min}\left(d(v,b)\right)\leq c\right\}$, then a $(S_{B},f)$-topology-aware strictly stabilizing protocol is a $(c,f)$-strictly stabilizing protocol. Then, the concept of topology-aware strict stabilization is a generalization of the strict stabilization. However, note that a TA strictly stabilizing protocol is stronger than a classical self-stabilizing protocol (that may never meet their specification in the presence of Byzantine processes). The parameter $S_{B}$ is introduced to quantify the strength of fault containment, we do not require each process to know the actual definition of the set.
Similarly to topology-aware strict stabilization, we can weaken the notion of strong stabilization using the notion of containment area. This idea was introduced by \cite{DMT10cd}. We recall in the following the formal definition of this concept.
\begin{definition}[$S_B$-stable configuration]
A configuration $\rho$ is $S_B$-stable if every $S_B$-correct process never changes the values of its O-variables as long as Byzantine processes make no action.
\end{definition}
\begin{definition}[$S_{B}$-TA-disruption]
A portion of execution $e=\rho_0,\rho_1,\ldots,\rho_t$ ($t>1$) is a $S_{B}$-TA-disruption if and only if the followings hold:
\begin{enumerate}
\item $e$ is finite,
\item $e$ contains at least one action of a $S_{B}$-correct process for changing the value of an O-variable,
\item $\rho_0$ is $S_{B}$-legitimate for $spec$ and $S_B$-stable, and
\item $\rho_t$ is the first configuration after $\rho_0$ such that $\rho_t$ is $S_{B}$-legitimate for $spec$ and $S_B$-stable.
\end{enumerate}
\end{definition}
\begin{definition}[$(t,k,S_{B},f)$-TA time contained configuration]
A configuration $\rho_0$ is $(t,k,S_{B},$ $f)$-TA time contained for \emph{spec} if given at most $f$ Byzantine processes, the following properties are satisfied:
\begin{enumerate}
\item $\rho_0$ is $S_{B}$-legitimate for \emph{spec} and $S_B$-stable,
\item every execution starting from $\rho_0$ contains a $S_B$-legitimate configuration for \emph{spec} after which the values of all the O-variables of $S_B$-correct processes remain unchanged (even when Byzantine processes make actions repeatedly and forever),
\item every execution starting from $\rho_0$ contains at most $t$ $S_B$-TA-disruptions, and
\item every execution starting from $\rho_0$ contains at most $k$ actions of changing the values of O-variables for each $S_B$-correct process.
\end{enumerate}
\end{definition}
\begin{definition}[$(t,S_{B},f)$-TA strongly stabilizing protocol]
A protocol $A$ is $(t,S_{B},f)$-TA\\ strongly stabilizing if and only if starting from any arbitrary configuration, every execution involving at most $f$ Byzantine processes contains a $(t,k,S_{B},f)$-TA-time contained configuration that is reached after at most $l$ actions of each $S_{B}$-correct node. Parameters $l$ and $k$ are respectively the $(t,S_{B},f)$-stabilization time and the $(t,S_{B},f)$-process-disruption time of $A$.
\end{definition}
\section{Maximum Metric Tree Construction}
\subsection{Definition and Specification}
In this work, we deal with maximum (routing) metric trees as defined in \cite{GS03j}. Informally, the goal of a routing protocol is to construct a tree that simultaneously maximizes the metric values of all of the nodes with respect to some total ordering $\prec$. In the following, we recall all definitions and notations introduced in \cite{GS03j}.
\begin{definition}[Routing metric]
A \emph{routing metric} (or just \emph{metric}) is a five-tuple $(M,W,met,mr,$ $\prec)$ where:
\begin{enumerate}
\item $M$ is a set of metric values,
\item $W$ is a set of edge weights,
\item $met$ is a metric function whose domain is $M\times W$ and whose range is $M$,
\item $mr$ is the maximum metric value in $M$ with respect to $\prec$ and is assigned to the root of the system,
\item $\prec$ is a less-than total order relation over $M$ that satisfies the following three conditions for arbitrary metric values $m$, $m'$, and $m''$ in $M$:
\begin{enumerate}
\item irreflexivity: $m\not\prec m$,
\item transitivity : if $m\prec m'$ and $m'\prec m''$ then $m\prec m''$,
\item totality: $m\prec m'$ or $m'\prec m$ or $m=m'$.
\end{enumerate}
\end{enumerate}
Any metric value $m\in M\setminus\{mr\}$ satisfies the \emph{utility condition} (that is, there exist $w_0,\ldots,w_{k-1}$ in $W$ and $m_0=mr,m_1,\ldots,m_{k-1},m_{k}=m$ in $M$ such that $\forall i\in\{1,\ldots,k\},m_i=met(m_{i-1},w_{i-1})$).
\end{definition}
For instance, we provide the definition of four classical metrics with this model: the shortest path metric ($\mathcal{SP}$), the flow metric ($\mathcal{F}$), and the reliability metric ($\mathcal{R}$). Note also that we can modelise the construction of a spanning tree with no particular constraints in this model using the metric $\mathcal{NC}$ described below and the construction of a BFS spanning tree using the shortest path metric ($\mathcal{SP}$) with $W_1=\{1\}$ (we denoted this metric by $\mathcal{BFS}$ in the following).
\[\begin{array}{rclrcl}
\mathcal{SP}&=&(M_1,W_1,met_1,mr_1,\prec_1)&\mathcal{F}&=&(M_2,W_2,met_2,mr_2,\prec_2)\\
\text{where}& & M_1=\mathbb{N}&\text{where}& & mr_2\in\mathbb{N}\\
&& W_1=\mathbb{N}&&& M_2=\{0,\ldots,mr_2\}\\
&& met_1(m,w)=m+w&&& W_2=\{0,\ldots,mr_2\}\\
&& mr_1=0&&& met_2(m,w)=min\{m,w\}\\
&& \prec_1 \text{ is the classical }>\text{ relation}&&& \prec_2 \text{ is the classical }<\text{ relation}
\end{array}\]
\[\begin{array}{rclrcl}
\mathcal{R}&=&(M_3,W_3,met_3,mr_3,\prec_3) & \mathcal{NC}&=&(M_4,W_4,met_4,mr_4,\prec_4)\\
\text{where}& & M_3=[0,1] & \text{where}& & M_4=\{0\} \\
&& W_3=[0,1] &&& W_4=\{0\}\\
&& met_3(m,w)=m*w &&& met_4(m,w)=0\\
&& mr_3=1 &&& mr_4=0\\
&& \prec_3 \text{ is the classical }<\text{ relation} &&& \prec_4 \text{ is the classical }<\text{ relation}
\end{array}\]
\begin{definition}[Assigned metric]
An \emph{assigned metric} over a system $S$ is a six-tuple $(M,W,met,$ $mr,\prec,wf)$ where $(M,W,met,mr,\prec)$ is a metric and $wf$ is a function that assigns to each edge of $S$ a weight in $W$.
\end{definition}
Let a rooted path (from $v$) be a simple path from a process $v$ to the root $r$. The next set of definitions are with respect to an assigned metric $(M,W,met,mr,\prec,wf)$ over a given system $S$.
\begin{definition}[Metric of a rooted path]
The \emph{metric of a rooted path} in $S$ is the prefix sum of $met$ over the edge weights in the path and $mr$.
\end{definition}
For example, if a rooted path $p$ in $S$ is $v_k,\ldots,v_0$ with $v_0=r$, then the metric of $p$ is $m_k=met(m_{k-1},wf(\{v_k,v_{k-1}\})$ with $\forall i\in\{1,\ldots,k-1\},m_i=met(m_{i-1},wf(\{v_i,v_{i-1}\})$ and $m_0=mr$.
\begin{definition}[Maximum metric path]
A rooted path $p$ from $v$ in $S$ is called a \emph{maximum metric path} with respect to an assigned metric if and only if for every other rooted path $q$ from $v$ in $S$, the metric of $p$ is greater than or equal to the metric of $q$ with respect to the total order $\prec$.
\end{definition}
\begin{definition}[Maximum metric of a node]
The \emph{maximum metric of a node} $v\neq r$ (or simply \emph{metric value} of $v$) in $S$ is defined by the metric of a maximum metric path from $v$. The maximum metric of $r$ is $mr$.
\end{definition}
\begin{definition}[Maximum metric tree]
A spanning tree $T$ of $S$ is a \emph{maximum metric tree} with respect to an assigned metric over $S$ if and only if every rooted path in $T$ is a maximum metric path in $S$ with respect to the assigned metric.
\end{definition}
The goal of the work of \cite{GS03j} is the study of metrics that always allow the construction of a maximum metric tree. More formally, the definition follows.
\begin{definition}[Maximizable metric]
A metric is \emph{maximizable} if and only if for any assignment of this metric over any system $S$, there is a maximum metric tree for $S$ with respect to the assigned metric.
\end{definition}
Given a maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$, the aim of this work is to study the construction of a maximum metric tree with respect to $\mathcal{M}$ which spans the system in a self-stabilizing way in a system subject to permanent Byzantine failures. It is obvious that these Byzantine processes may disturb some correct processes. It is why, we relax the problem in the following way: we want to construct a maximum metric forest with respect to $\mathcal{M}$. The root of any tree of this forest must be either the real root or a Byzantine process.
Each process $v$ has two O-variables: a pointer to its parent in its tree ($prnt_v\in N_v\cup\{\bot\}$) and a level which stores its current metric value ($level_v\in M$). Obviously, Byzantine process may disturb (at least) their neighbors. We use the following specification of the problem.
We introduce new notations as follows. Given an assigned metric $(M,W,met,mr,\prec,wf)$ over the system $S$ and two processes $u$ and $v$, we denote by $\mu(u,v)$ the maximum metric of node $u$ when $v$ plays the role of the root of the system. If $u$ and $v$ are neighbors, we denote by $w_{u,v}$ the weight of the edge $\{u,v\}$ (that is, the value of $wf(\{u,v\})$).
\begin{definition}[$\mathcal{M}$-path]
Given an assigned metric $\mathcal{M}=(M,W,mr,met,\prec,wf)$ over a system $S$, a path $(v_0,\ldots,v_k)$ ($k\geq 1$) of $S$ is a \emph{$\mathcal{M}$-path} if and only if:
\begin{enumerate}
\item $prnt_{v_0}=\bot$, $level_{v_0}=0$, and $v_0\in B\cup\{r\}$,
\item $\forall i\in\{1,\ldots,k\}, prnt_{v_i}=v_{i-1}$ and $level_{v_i}=met(level_{v_{i-1}},w_{v_i,v_{i-1}})$,
\item $\forall i\in\{1,\ldots,k\}, met(level_{v_{i-1}},w_{v_i,v_{i-1}})=\underset{u\in N_v}{max_\prec}\{met(level_{u},w_{v_i,u})\}$, and
\item $level_{v_{k}}=\mu(v_k,v_0)$.
\end{enumerate}
\end{definition}
We define the specification predicate $spec(v)$ of the maximum metric tree construction with respect to a maximizable metric $\mathcal{M}$ as follows.
\[spec(v) : \begin{cases}
prnt_v = \bot \text{ and } level_v = 0 \text{ if } v \text{ is the root } r \\
\text{there exists a }\mathcal{M}\text{-path } (v_0,\ldots,v_k) \text{ such that } v_k=v \text{ otherwise}
\end{cases}\]
\subsection{Previous results}
In this section, we summarize known results about maximum metric tree construction. The first interesting result about maximizable metrics is due to \cite{GS03j} that provides a fully characterization of maximizable metrics as follow.
\begin{definition}[Boundedness]
A metric $(M,W,met,mr,\prec)$ is \emph{bounded} if and only if: $\forall m \in M,\forall w\in W, met(m,w)\prec m \text{ or }met(m,w)=m$
\end{definition}
\begin{definition}[Monotonicity]
A metric $(M,W,met,mr,\prec)$ is \emph{monotonic} if and only if: $\forall (m,$ $m')\in M^2,\forall w\in W, m\prec m'\Rightarrow (met(m,w)\prec met(m',w)\text{ or }met(m,w)=met(m',w))$
\end{definition}
\begin{theorem}[Characterization of maximizable metrics \cite{GS03j}]
A metric is maximizable if and only if this metric is bounded and monotonic.
\end{theorem}
Secondly, \cite{GS99c} provides a self-stabilizing protocol to construct a maximum metric tree with respect to any maximizable metric. Now, we focus on self-stabilizating solutions resilient to Byzantine faults. Following discussion of Section 2, it is obvious that there exists no strictly stabilizing protocol for this problem. If we consider the weaker notion of topology-aware strict stabilization, \cite{DMT10ca} defines the best containment area as:
\[S_{B}=\left\{v\in V\setminus B\left|\mu(v,r)\preceq max_\prec\{\mu(v,b),b\in B\}\right.\right\}\setminus\{r\}\]
Intuitively, $S_B$ gathers correct processes that are closer (or at equal distance) from a Byzantine process than the root according to the metric. Moreover, \cite{DMT10ca} proves that the algorithm introduced for the maximum metric spanning tree construction in \cite{GS99c} performed this optimal containment area. More formally, \cite{DMT10ca} proves the following results.
\begin{theorem}[\cite{DMT10ca}]\label{th:impTAstrict}
Given a maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$, even under the central daemon, there exists no $(A_B,1)$-TA-strictly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{M}$ where $A_B\varsubsetneq S_B$.
\end{theorem}
\begin{theorem}[\cite{DMT10ca}]\label{th:SSMAXstrict}
Given a maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$, the protocol of \cite{GS99c} is a $(S_B,n-1)$-TA strictly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{M}$.
\end{theorem}
Some others works try to circumvent the impossibility result of strict stabilization using the concept ot strong stabilization but do not provide results for any maximizable metric. Indeed, \cite{DMT11j} proves the following result about spanning tree.
\begin{theorem}[\cite{DMT11j}]
There exists a $(t,0,n-1)$-strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{NC}$ (that is, for a spanning tree with no particular constraints) with a finite $t$.
\end{theorem}
On the other hand, regarding BFS spanning tree construction, \cite{DMT10cd} proved the following impossibility result.
\begin{theorem}[\cite{DMT10cd}]
Even under the central daemon, there exists no $(t,c,1)$-strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{BFS}$ where $t$ and $c$ are two finite integers.
\end{theorem}
These two results motivate our result related to strong stabilization in the general case (see Section 4.1) that proves a necessary condition on the maximizable metric to allow strong stabilization.
Now, if we focus on topology-aware strong stabilization, \cite{DMT10cd} proved the following results.
\begin{theorem}[\cite{DMT10cd}]\label{th:impTAStrongBFS}
Even under the central daemon, there exists no $(t,A_B^*,1)$-TA strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{BFS}$ where $A_B^*\varsubsetneq \{v\in V|\underset{b\in B}{min}(d(v,b))<d(r,v)\}$ and $t$ is a finite integer.
\end{theorem}
\begin{theorem}[\cite{DMT10cd}]
The protocol of \cite{HC92j} is a $(t,S_B^*,n-1)$-TA strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{BFS}$ where $t$ is a finite integer and $S_B^*=\{v\in V|\underset{b\in B}{min}(d(v,b))<d(r,v)\}$.
\end{theorem}
In the following, we generalize the Theorem \ref{th:impTAStrongBFS} to any maximizable metric (see Section 4.2).
\section{Necessary conditions}
In this section, we provide our necessary conditions about containment radius (respectively area) of any strongly stabilizing (respectively TA strongly stabilizing) protocol for the maximum metric tree construction.
\subsection{Strong Stabilization}
We introduce here some new definitions to characterize some important properties of maximizable metrics that are used in the following.
\begin{definition}[Strictly decreasing metric]
A metric $\mathcal{M}=(M,W,mr,met,\prec)$ is \emph{strictly decreasing} if, for any metric value $m\in M$, the following property holds: either $\forall w\in W,met(m,w)\prec m$ or $\forall w\in W,met(m,w)=m$.
\end{definition}
\begin{definition}[Fixed point]
A metric value $m$ is a \emph{fixed point} of a metric $\mathcal{M}=(M,W,mr,met,\prec)$ if $m\in M$ and if for any value $w\in W$, we have: $met(m,w)=m$.
\end{definition}
Then, we define a specific class of maximizable metrics and we prove that it is possible to construct a maximum metric tree in a strongly-stabilizing way only if we consider such a metric.
\begin{definition}[Strongly maximizable metric]
A maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$ is strongly maximizable if and only if $|M|=1$ or if the following properties holds:
\begin{itemize}
\item $|M|\geq 2$,
\item $\mathcal{M}$ is strictly decreasing, and
\item $\mathcal{M}$ has one and only one fixed point.
\end{itemize}
\end{definition}
Note that $\mathcal{NC}$ is a strongly maximizable metric (since $|M_4|=1$) whereas $\mathcal{BFS}$ or $\mathcal{SP}$ are not (since the first one has no fixed point, the second is not strictly decreasing). If we consider the metric $\mathcal{MET}$ defined below, we can show that $\mathcal{MET}$ is a strongly maximizable metric such that $|M|\geq 2$.
\[\begin{array}{rcl}
\mathcal{MET}&=&(M_5,W_5,met_5,mr_5,\prec_5)\\
\text{where}& & M_5=\{0,1,2,3\}\\
&& W_5=\{1\}\\
&& met_5(m,w)=max\{0,m-w\}\\
&& mr_5=3\\
&& \prec_5 \text{ is the classical }<\text{ relation}
\end{array}\]
Now, we can state our first necessary condition.
\begin{theorem}\label{th:necessarConditionStrong}
Given a maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$, even under the central daemon, there exists no $(t,c,1)$-strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{M}$ for any finite integer $t$ if:
\[\left\{\begin{array}{l}
\mathcal{M} \mbox{ is not a strongly maximizable metric}\\
\mbox{ or }\\
c<|M|-2
\end{array}\right.\]
\end{theorem}
\begin{proof}
We prove this result by contradiction. We assume so that $\mathcal{M}=(M,W,mr,met,\prec)$ is a maximizable metric such that there exist a finite integer $t$ and a protocol $\mathcal{P}$ that is a $(t,c,1)$-strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{M}$. We distinguish the following cases (note that they are exhaustive):
\begin{description}
\item[Case 1:] $\mathcal{M}$ is a strongly maximizing metric and $c<|M|-2$.
As $c\geq 0$, we know that $|M|\geq 2$ and by definition of a strongly stabilizing metric, $\mathcal{M}$ is strictly decreasing, and $\mathcal{M}$ has one and only one fixed point.
\begin{figure}[t]
\noindent \begin{centering} \include{possStrongCase1}
\par\end{centering}
\caption{Configurations used in proof of Theorem \ref{th:necessarConditionStrong}, case 1.}
\label{fig:possStrongCase1}
\end{figure}
By assumption on $\mathcal{M}$, we know that there exist $c+3$ distinct metric values $m_0=mr,m_1,\ldots,$ $m_{c+2}$ in $M$ and $w_0,w_1,\ldots,w_{c+1}$ in $W$ such that: $\forall i\in\{1,\ldots,c+2\},m_i=met(m_{i-1},w_{i-1})\prec m_{i-1}$.
Let $S=(V,E,\mathcal{W})$ be the following weighted system $V=\{p_0=r,p_1,\ldots,p_{2c+2},p_{2c+3}=b\}$, $E=\{\{p_i,p_{i+1}\},i\in\{0,\ldots,2c+2\}\}$ and $\forall i\in\{0,c+1\},w_{p_i,p_{i+1}}=w_{p_{2c+3-i},p_{2c+2-i}}=w_i$. Note that the choice $w_{p_{c+1},p_{c+2}}=w_{c+1}$ ensures us the following property when $level_r=level_b=mr$: $\mu(p_{c+1},b)\prec\mu(p_{c+1},r)$ (and by symmetry, $\mu(p_{c+2},r)\prec\mu(p_{c+2},b)$). Process $p_0$ is the real root and process $b$ is a Byzantine one. Note that the construction of $\mathcal{W}$ ensures the following properties when $level_r=level_b=mr$: $\forall i\in\{1,\ldots,c+1\},\mu(p_i,r)=\mu(p_{2c+3-i},b)$, $\mu(p_i,b)\prec\mu(p_i,r)$ and $\mu(p_{2c+3-i},r)\prec\mu(p_{2c+3-i},b)$.
Assume that the initial configuration $\rho_0$ of $S$ satisfies: $prnt_r=prnt_b=\bot$, $level_r=level_b=mr$, and other variables of $b$ (if any) are identical to those of $r$ (see Figure \ref{fig:possStrongCase1}, variables of other processes may be arbitrary). Assume now that $b$ takes exactly the same actions as $r$ (if any) immediately after $r$. Then, by symmetry of the execution and by convergence of $\mathcal{P}$ to $spec$, we can deduce that the system reaches in a finite time a configuration $\rho_1$ (see Figure \ref{fig:possStrongCase1}) in which: $\forall i\in\{1,\ldots,c+1\}, prnt_{p_i}=p_{i-1}$, $level_{p_i}=\mu(p_i,r)=m_i$ and $\forall i\in\{c+2,\ldots,2c+2\},prnt_{p_i}=p_{i+1}$ and $level_{p_i}=\mu(p_{i},b)=m_{2c+3-i}$ (because this configuration is the only one in which all correct process $v$ satisfies $spec(v)$ when $prnt_r=prnt_b=\bot$ and $level_r=level_b=mr$ by construction of $\mathcal{W}$). Note that $\rho_1$ is $c$-legitimate and $c$-stable.
Assume now that the Byzantine process acts as a correct process and executes correctly its algorithm. Then, by convergence of $\mathcal{P}$ in fault-free systems (remember that a strongly-stabilizing algorithm is a special case of self-stabilizing algorithm), we can deduce that the system reach in a finite time a configuration $\rho_2$ (see Figure \ref{fig:possStrongCase1}) in which: $\forall i\in\{1,\ldots,2c+3\},prnt_{p_i}=p_{i-1}$ and $level_{p_i}=\mu(p_i,r)$ (because this configuration is the only one in which all process $v$ satisfies $spec(v)$). Note that the portion of execution between $\rho_1$ and $\rho_2$ contains at least one $c$-perturbation ($p_{c+2}$ is a $c$-correct process and modifies at least once its O-variables) and that $\rho_2$ is $c$-legitimate and $c$-stable.
Assume now that the Byzantine process $b$ takes the following state: $prnt_{b}=\bot$ and $level_b=mr$. This step brings the system into configuration $\rho_3$ (see Figure \ref{fig:possStrongCase1}). From this configuration, we can repeat the execution we constructed from $\rho_0$. By the same token, we obtain an execution of $\mathcal{P}$ which contains $c$-legitimate and $c$-stable configurations (see $\rho_1$) and an infinite number of $c$-perturbation which contradicts the $(t,c,1)$-strong stabilization of $\mathcal{P}$.
\item[Case 2:] $\mathcal{M}$ is not strictly decreasing.
\begin{figure}[t]
\noindent \begin{centering} \include{possStrongCase23}
\par\end{centering}
\caption{Configurations used in proof of Theorem \ref{th:necessarConditionStrong}, cases 2 and 3.}
\label{fig:possStrongCase23}
\end{figure}
By definition, we know that $\mathcal{M}$ is not a strongly maximizable metric. Hence, we have $|M|\geq 2$. Then, the definition of a strictly decreasing metric implies that there exists a metric value $m\in M$ such that: $\exists w\in W,$ $met(m,w)=m$ and $\exists w'\in W,m'=met(m,w')\prec m$ (and thus $m$ is not a fixed point of $\mathcal{M}$). By the utility condition on $M$, we know that there exists a sequence of metric values $m_0=mr,m_1,\ldots,m_l=m$ in $M$ and $w_0,w_1,\ldots,w_{l-1}$ in $W$ such that $\forall i\in\{1,\ldots,l\},m_i=met(m_{i-1},w_{i-1})$. Denote by $k$ the length of the shortest such sequence. Note that this implies that $\forall i\in\{1,\ldots,k\},m_i\prec m_{i-1}$ (otherwise we can remove $m_i$ from the sequence and this is contradictory with the construction of $k$). We distinguish the following cases:
\begin{description}
\item[Case 2.1:] $k\geq c+2$.\\
We can use the same token as case 1 above by using $w'$ instead of $w_{c+1}$ in the case where $k=c+2$ (since we know that $met(m,w')\prec m$).
\item[Case 2.2:] $k< c+2$.\\
Let $S_1=(V,E,\mathcal{W})$ be the following weighted system $V=\{p_0=r,p_1,\ldots,p_{2c+2},p_{2c+3}=b\}$, $E=\{\{p_i,p_{i+1}\},i\in\{0,\ldots,2c+2\}\}$, $\forall i\in\{0,\ldots,k-1\},w_{p_i,p_{i+1}}=w_{p_{2c+3-i},p_{2c+2-i}}=w_i$, $\forall i\in\{k,\ldots,c\},w_{p_i,p_{i+1}}=w_{p_{2c+3-i},p_{2c+2-i}}=w$ and $w_{p_{c+1},p_{c+2}}=w'$ (see Figure \ref{fig:possStrongCase23}). Note that this choice ensures us the following property when $level_r=level_b=mr$: $\mu(p_{c+1},b)\prec\mu(p_{c+1},r)$ (and by symmetry, $\mu(p_{c+2},r)\prec\mu(p_{c+2},b)$). Process $p_0$ is the real root and process $b$ is a Byzantine one. Note that the construction of $\mathcal{W}$ ensures the following properties when $level_r=level_b=mr$: $\forall i\in\{1,\ldots,c+1\},\mu(p_i,r)=\mu(p_{2c+3-i},b)$, $\mu(p_i,b)\prec\mu(p_i,r)$ and $\mu(p_{2c+3-i},r)\prec\mu(p_{2c+3-i},b)$.
This construction allows us to follow the same proof as in case 1 above.
\end{description}
\item[Case 3:] $\mathcal{M}$ has no or more than two fixed point, and is strictly decreasing.
If $\mathcal{M}$ has no fixed point and is strictly decreasing, then $|M|$ is not finite and then, we can apply the result of case 1 above since $c$ is a finite integer.
If $\mathcal{M}$ has two or more fixed points and is strictly decreasing, denote by $\Upsilon$ and $\Upsilon'$ two fixed points of $\mathcal{M}$. Without loss of generality, assume that $\Upsilon\prec\Upsilon'$. By the utility condition on $M$, we know that there exists sequences of metric values $m_0=mr,m_1,\ldots,m_l=\Upsilon$ and $m'_0=mr,m'_1,\ldots,m'_{l'}=\Upsilon'$ in $M$ and $w_0,w_1,\ldots,w_{l-1}$ and $w'_0,w'_1,\ldots,w'_{l'-1}$ in $W$ such that $\forall i\in\{1,\ldots,l\},m_i=met(m_{i-1},w_{i-1})$ and $\forall i\in\{1,\ldots,l'\},m'_i=met(m'_{i-1},w'_{i-1})$. Denote by $k$ and $k'$ the length of shortest such sequences. Note that this implies that $\forall i\in\{1,\ldots,k\},m_i\prec m_{i-1}$ and $\forall i\in\{1,\ldots,k'\},m'_i\prec m'_{i-1}$ (otherwise we can remove $m_i$ or $m'_i$ from the corresponding sequence). We distinguish the following cases:
\begin{description}
\item[Case 3.1:] $k>c+2$ or $k'>c+2$.\\
Without loss of generality, assume that $k>c+2$ (the second case is similar). We can use the same token as case 1 above.
\item[Case 3.2:] $k\leq c+2$ and $k'\leq c+2$.\\
Let $w$ be an arbitrary value of $W$. Let $S_2=(V,E,\mathcal{W})$ be the following weighted system $V=\{p_0=r,p_1,\ldots,p_{2c+2},p_{2c+3}=b\}$, $E=\{\{p_i,p_{i+1}\},i\in\{0,\ldots,2c+2\}\}$, $\forall i\in\{0,k-1\},w_{p_i,p_{i+1}}=w_i$, $\forall i\in\{0,k'-1\},w_{p_{2c+3-i},p_{2c+2-i}}=w'_i$ and $\forall i\in\{k,2c+2-k'\}, w_{p_{i},p_{i+1}}=w$ (see Figure \ref{fig:possStrongCase23}). Note that this choice ensures us the following property when $level_r=level_b=mr$: $\mu(p_{c+1},r)=\Upsilon\prec\Upsilon' =\mu(p_{c+1},b)$ and $\mu(p_{c+2},r)=\Upsilon\prec\Upsilon'=\mu(p_{c+2},b)$. Process $p_0$ is the real root and process $b$ is a Byzantine one.
This construction allows us to follow a similar proof as in case 1 above (note that any process $u$ which satisfies $\mu(u,r)\prec\Upsilon'$ will be disturb infinitely often, in particular at least $p_{c+1}$ and $p_{c+2}$ which contradicts the $(t,c,1)$-strong stabilization of $\mathcal{P}$).
\end{description}
\end{description}
In any case, we show that there exists a system which contradicts the $(t,c,1)$-strong stabilization of $\mathcal{P}$ that ends the proof.
\end{proof}
\subsection{Topology Aware Strong Stabilization}
First, we generalize the set $S_B^*$ previously defined for the $\mathcal{BFS}$ metric in \cite{DMT10cd} to any maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$.
\[S_{B}^*=\left\{v\in V\setminus B\left|\mu(v,r)\prec\underset{b\in B}{max_\prec}\{\mu(v,b)\}\right.\right\}\]
Intuitively, $S_B^*$ gathers the set of corrects processes that are strictly closer (according to $\mathcal{M}$) to a Byzantine process than the root. Figures from \ref{fig:ExSP} to \ref{fig:ExReliability} provide some examples of containment areas with respect to several maximizable metrics and compare it to $S_B$, the optimal containment area for TA strict stabilization.
\begin{figure}[t]
\noindent \begin{centering} \include{ExSP}
\par\end{centering}
\caption{Examples of containment areas for SP spanning tree construction.}
\label{fig:ExSP}
\end{figure}
\begin{figure}[t]
\noindent \begin{centering} \include{ExFlow}
\par\end{centering}
\caption{Examples of containment areas for flow spanning tree construction.}
\label{fig:ExFlow}
\end{figure}
\begin{figure}[t]
\noindent \begin{centering} \include{ExReliability}
\par\end{centering}
\caption{Examples of containment areas for reliability spanning tree construction.}
\label{fig:ExReliability}
\end{figure}
Now, we can state our generalization of Theorem \ref{th:impTAStrongBFS}.
\begin{theorem}\label{th:impTAstrong}
Given a maximizable metric $\mathcal{M}=(M,W,mr,met,\prec)$, even under the central daemon, there exists no $(t,A_B^*,1)$-TA-strongly stabilizing protocol for maximum metric spanning tree construction with respect to $\mathcal{M}$ where $A_B^*\varsubsetneq S_B^*$ and $t$ is a given finite integer.
\end{theorem}
\begin{proof}
Let $\mathcal{M}=(M,W,mr,met,\prec)$ be a maximizable metric and $\mathcal{P}$ be a $(t,A_B^*,1)$-TA-strongly stabilizing protocol for maximum metric spanning tree construction protocol with respect to $\mathcal{M}$ where $A_B^*\varsubsetneq S_B^*$ and $t$ is a finite integer. We must distinguish the following cases:
\begin{description}
\item[Case 1:] $|M|=1$.\\
Denote by $m$ the metric value such that $M=\{m\}$. For any system and for any process $v$, we have $\mu(v,r)=\underset{b\in B}{min_\prec}\{\mu(v,b)\}=m$. Consequently, $S_B^*=\emptyset$ for any system. Then, it is absurd to have $A_B^*\varsubsetneq S_B^*$.
\item[Case 2:] $|M|\geq 2$.\\
By definition of a bounded metric, we can deduce that there exists $m\in M$ and $w\in W$ such that $m=met(mr,w)\prec mr$. Then, we must distinguish the following cases:
\begin{description}
\item[Case 2.1:] $m$ is a fixed point of $\mathcal{M}$.\\
Let $S$ be a system such that any edge incident to the root or a Byzantine process has a weight equals to $w$. Then, we can deduce that we have: $m=\underset{b\in B}{max_\prec}\{\mu(r,b)\}\prec \mu(r,r)=mr$ and for any correct process $v\neq r$, $\mu(v,r)=\underset{b\in B}{max_\prec}\{\mu(v,b)\}=m$. Hence, $S_B^*=\emptyset$ for any such system. Then, it is absurd to have $A_B^*\varsubsetneq S_B^*$.
\item[Case 2.2:] $m$ is not a fixed point of $\mathcal{M}$.\\
This implies that there exists $w'\in W$ such that: $met(m,w')\prec m$ (remember that $\mathcal{M}$ is bounded). Consider the following system: $V=\{r,u,u',v,v',b\}$, $E=\{\{r,u\},\{r,u'\},$ $\{u,v\},\{u',v'\},\{v,b\},\{v',b\}\}$, $w_{r,u}=w_{r,u'}=w_{v,b}=w_{v',b}=w$, and $w_{u,v}=w_{u',v'}=w'$ ($b$ is a Byzantine process). We can see that $S_B^*=\{v,v'\}$. Since $A_B^*\varsubsetneq S_B$, we have: $v\notin A_B^*$ or $v'\notin A_B^*$. Consider now the following configuration $\rho_0$: $prnt_r=prnt_b=\bot$, $level_r=level_b=mr$, and $prnt$, $level$ variables of other processes are arbitrary (see Figure \ref{fig:impTAstrong}, other variables may have arbitrary values but other variables of $b$ are identical to those of $r$).
Assume now that $b$ takes exactly the same actions as $r$ (if any) immediately after $r$ (note that $r\notin A_B^*$ and hence $prnt_r=\bot$ and $level_r=mr$ still hold by closure and then $prnt_b=\bot$ and $level_b=mr$ still hold too). Then, by symmetry of the execution and by convergence of $\mathcal{P}$ to $spec$, we can deduce that the system reaches in a finite time a configuration $\rho_1$ (see Figure \ref{fig:impTAstrong}) in which: $prnt_r=prnt_b=\bot$, $prnt_u=prnt_{u'}=r$, $prnt_v=prnt_{v'}=b$, $level_r=level_b=mr$, and $level_u=level_{u'}=level_v=level_{v'}=m$ (because this configuration is the only one in which all correct process $v$ satisfies $spec(v)$ when $prnt_r=prnt_b=\bot$ and $level_r=level_b=mr$ since $met(m,w')\prec m$). Note that $\rho_1$ is $A_B^*$-legitimate for $spec$ and $A_B^*$-stable (whatever $A_B^*$ is).
Assume now that $b$ behaves as a correct processor with respect to $\mathcal{P}$. Then, by convergence of $\mathcal{P}$ in a fault-free system starting from $\rho_1$ which is not legitimate (remember that a TA-strongly stabilizing algorithm is a special case of self-stabilizing algorithm), we can deduce that the system reach in a finite time a configuration $\rho_2$ (see Figure \ref{fig:impTAstrong}) in which: $prnt_r=\bot$, $prnt_u=prnt_{u'}=r$, $prnt_v=u$, $prnt_{v'}=u'$, $prnt_b=v$ (or $prnt_b=v'$), $level_r=mr$, $level_u=level_{u'}=m$ $level_v=level_{v'}=met(m,w')=m'$, and $level_b=met(m',w)=m''$. Note that processes $v$ and $v'$ modify their O-variables in the portion of execution between $\rho_1$ and $\rho_2$ and that $\rho_2$ is $A_B^*$-legitimate for $spec$ and $A_B^*$-stable (whatever $A_B^*$ is). Consequently, this portion of execution contains at least one $A_B^*$-TA-disruption (whatever $A_B^*$ is).
Assume now that the Byzantine process $b$ takes the following state: $prnt_b=\bot$ and $level_b=mr$. This step brings the system into configuration $\rho_3$ (see Figure \ref{fig:impTAstrong}). From this configuration, we can repeat the execution we constructed from $\rho_0$. By the same token, we obtain an execution of $\mathcal{P}$ which contains $c$-legitimate and $c$-stable configurations (see $\rho_1$) and an infinite number of $A_B^*$-TA-disruption (whatever $A_B^*$ is) which contradicts the $(t,A_B^*,1)$-TA-strong stabilization of $\mathcal{P}$.
\end{description}
\end{description}
\end{proof}
\begin{figure}[t]
\noindent \begin{centering} \include{impTAstrong}
\par\end{centering}
\caption{Configurations used in proof of Theorem \ref{th:impTAstrong}.}
\label{fig:impTAstrong}
\end{figure}
\section{Conclusion}
In this paper, we presented two necessary conditions to achieve strong stabilization and topology-aware strong stabilization in maximum metric tree construction. Our work obviously leads to the following open question: is there a topology-aware strongly stabilizing protocol that ensures a containmemt area equal to $S_B^*$? We conjecture that it is the case.
\bibliographystyle{plain}
|
\section{Introduction}
Study of diffuse background emission produced by faint sources with flux levels below the sensitivity of a telescope is commonly used to constrain the nature of source populations in the Universe and their cosmological evolution. In the high-energy $\gamma$-ray\ band (HE, 0.1-100 GeV) the diffuse extragalactic $\gamma$-ray\ background (EGB) was detected for the first time by {\it SAS-2} satellite \citep{fichtel78}, was further studied by EGRET telescope on board of {\it CGRO} mission \citep{sreekumar98,strong04} and, most recently, by the Large Area Telescope (LAT) on board of {\it Fermi} satellite \citep{fermi_background}.
It is often assumed that the dominant contribution to the EGB is given by distant Active Galactic Nuclei (AGN), in particular, blazars \citep{padovani93,stecker93,chiang95,stecker96,mukherjee99,mucke00,inoue09}. However, a recent study by {\it Fermi} collaboration reveals that blazars might contribute only a relatively small fraction of the HE EGB level \citep{fermi_agn}, while a significant part of the EGB should be either explained by a yet unknown source population or have a truly diffuse nature (see, however, \citet{stecker10}).
The EGB in the Very-High-Energy (VHE, $\gamma$-ray s in the $E\gtrsim 100$~GeV) range has never been measured. On one hand, the effective collection area of previous space-based $\gamma$-ray\ telescopes was not sufficient to achieve significant photon statistics in this energy band. On the other hand, the efficiency of cosmic ray background rejection in the ground-based Cherenkov $\gamma$-ray\ telescopes, like HESS, MAGIC and VERITAS is not sufficient for detection of the isotropic diffuse EGB on top of the cosmic ray background. Thus, the properties and the origin of VHE EGB remain largely unconstrained up to now.
It is clear that the VHE EGB should contain a contribution from the unresolved point sources. The main candidate source class is, as in the case of HE EGB, that of blazars. At the same time, the VHE EGB could contain, apart from the contribution from unresolved extragalactic point sources, genuine diffuse components which could be produced via several mechanisms. For example, if the spectra of a large number of $\gamma$-ray -loud AGNs extend to the energies above 300~GeV, all the power emitted initially in
$\gamma$-ray s with energies higher than $\sim 300$~GeV is absorbed in the pair
production of $\gamma$-ray s on the cosmological infrared and/or microwave backgrounds \citep{gould67}.
Secondary inverse Compton emission of electron-positron pairs deposited in the
intergalactic space in result of the pair production leads to generation of
diffuse extragalactic emission in the VHE energy band \citep{coppi97}. Another mechanism
which can lead to the generation of diffuse component of VHE EGB is electromagnetic cascade initiated in the intergalactic space by ultra-high energy cosmic rays (UHECR) interacting with cosmic microwave background
photons \citep{berezinsky75,semikoz09,berezinsky11}. The cascade channels the power from the highest energies of about
$10^{20}$~eV down to the $\sim 0.1$~GeV band in which the mean free path of the
$\gamma$-ray s becomes comparable to the size of the visible part of the Universe.
Apart from the "guaranteed" (but, possibly, very weak) diffuse contributions, isotropic VHE $\gamma$-ray\ background might contain contributions from "exotic" diffuse sources, like diffuse emission from annihilation of Dark Matter particles in the outer halo of the Milky Way galaxy and the annihilation signal accumulated from the dark matter halos of all galaxies in the course of cosmological evolution \citep{fermi_DM}.
Whatever are the sources of VHE EGB, they are scattered across the Universe, so that a significant contribution to the flux is produced at redshifts $z\sim 1$. A known effect of absorption of VHE $\gamma$-ray s due to the interactions with infrared/optical Extragalactic Background Light (EBL) should lead to attenuation of the $E>50$~GeV signal produced by the sources at large redshifts $z\sim 1$ \citep{gould67,kneiske04,franceschini08,stecker_ebl,gilmore09}. This should leave an "imprint" on the VHE EBL spectrum, which should have the form of a gradual suppression with the increasing photon energy. Detecting a EBL suppression feature in the EGB spectrum would provide an important constraint on the (largely uncertain) evolution of the EBL density and spectrum up to redshifts $z\sim 1$. Such a constraint is otherwise difficult to obtain from the studies of individual extragalactic VHE $\gamma$-ray\ sources because of the limited signal statistics at the highest energies, especially for the sources at significant redshifts.
In what follows we discuss the measurement of the EGB derived from the data of {\it Fermi}/LAT telescope \citep{fermi_description}. The measurement is obtained from the counting of photons at high Galactic latitudes, after subtraction of the Galactic diffuse emission and the residual cosmic-ray background not rejected by the LAT data analysis software. We compare the measurement of EGB obtained in this way with the measurement previously derived from the likelihood analysis of all-sky data by \citet{fermi_background}.
The EGB flux above 30~GeV turns out to be comparable to the flux in extragalactic VHE $\gamma$-ray\ sources resolved by {\it Fermi}. We find that the spectrum of EGB in this energy range follows the cumulative spectrum of the resolved sources.
Dominant population of extragalactic VHE $\gamma$-ray\ sources is BL Lacs. Noticing the similarity of the spectrum of VHE EGB and of the cumulative BL Lac VHE $\gamma$-ray\ spectrum, we put forward a hypothesis that the VHE EBL is produced by unresolved BL Lacs with fluxes below the sensitivity of LAT. We explore this hypothesis and show that it could be valid if BL Lacs follow a positive cosmological evolution pattern, characteristic for other types of AGN, in particular for the parent population of BL Lacs objects, Fanaroff-Riley type I (FR I) radio galaxies.
\section{Data selection and data analysis}
For our analysis we consider all publicly available LAT data from August 4, 2008 to January 23, 2011. We process the data using {\it Fermi} Science Tools\footnote{\tt http://fermi.gsfc.nasa.gov/ssc/data/analysis/scitools/}. We filter the entire data set with {\it gtselect} and {\it gtmktime} tools following the recommendations of {\it Fermi} team\footnote{\tt http://fermi.gsfc.nasa.gov/ssc/data/analysis/scitools/} and retain only events belonging to "ultraclean" ({\tt P7ULTRACLEAN\_V6}) event class, which has minimal residual cosmic ray contamination.
Estimate of the contribution of point sources to the total flux requires separation of the photons coming from the point sources from those produced by the diffuse emission. Such separation is most straightforward for the photons with narrow point-spread-function (PSF). Taking this into account, select two sub-classes (which provide dominant contribution to the ultraclean events) with the most compact PSF, the sub-classes selected by imposing the selection criterium {\tt EVENT\_CLASS=65311} or {\tt 32543}. Other sub-classes of the ultraclean events have worse PSF. Point source contribution in these photons suffers from an additional uncertainty. Taking this into account, we restrict our attention to the subset of the ultraclean events with the best PSF.
We retain events with Earth zenith angle $\theta_z\le 100^\circ$. To estimate the flux from the photon counts we use {\it gtexposure} tool. We consider only events at high Galactic latitudes, in the regions $|b|\ge 60^\circ$.
Our analysis is based on the so-called "Pass 7" selection of the LAT data (see {\tt http://fermi.gsfc.nasa.gov/ssc/data/access/}). However, use use a comparison of the Pass 7 data with the previous Pass 6 data selection in the estimate residual cosmic ray contamination of the set of events chosen for the analysis. The residual cosmic ray fraction in the Pass 6 data was studied in details by \citet{fermi_background}. Re-calculation of the residual cosmic ray fraction for any new selection of events, including the one considered in our analysis, could be done in a straightforward way as explained in Section \ref{sec:cr}.
\section{Diffuse $\gamma$-ray background}
Signal detected by LAT at high Galactic latitudes contains four types of contributions: emission from point sources, diffuse $\gamma$-ray\ emission from the Galaxy, EGB and residual cosmic ray background not rejected by the analysis software. To measure the EGB flux, one needs to separate the contributions from the four components in the overall signal in a given energy band.
\subsection{Point source contribution}
\label{sec:ps}
Point source component could be singled out in a straightforward way if the set of sources detectable in a given energy band is known. To define the set of sources we find the sources correlating with the arrival directions of photons in each energy band, using the method described by \citet{100GeV_sky}\footnote{See {\tt http:/www.isdc.unige.ch/vhe/index.html} for an updated version of the VHE source list.}. To calculate the total number of photons associated to the sources, we construct a cumulative distribution of photons as a function of the distance $\theta$ from the source and split it on the background and source contributions. The background contribution grows asymptotically as $\theta^2$, while the source contribution asymptotically reaches constant. An example of the cumulative photon distribution around the source positions in the 12.5-25 GeV energy band is shown in Fig. \ref{fig:PSF_cumulative}.
\begin{figure}
\includegraphics[height=\linewidth,angle=-90]{fig1}
\caption{Cumulative front photon distributions around point sources in 12.5-25 GeV energy band. Red points show the source photons, blue points show the background. Horizontal lines (from top to bottom) show the 100\%, 95\% and 68\% levels.}
\label{fig:PSF_cumulative}
\end{figure}
\subsection{Galactic diffuse emission contribution}
\label{sec:gal}
Contribution of the diffuse emission from the Galaxy should be found via a detailed fitting the all-sky photon distribution to an all-sky spatial and spectral template. This contribution is best constrained by the all-sky photon distribution in the 0.1-10~GeV energy band, where event statistics is very high. Detailed fitting of the Galactic diffuse emission to the data in the 0.1-10~GeV band was done by \citet{fermi_background}. In our analysis we rely on the best-fit model of Galactic diffuse emission derived by \citet{fermi_background}. This model is available in the sky region of interest, $|b|\ge 60^\circ$, see Fig. 6 of the Supplemental Material in the Ref. \citet{fermi_background}. The uncertainties of the Galactic diffuse emission model are also discussed by \citet{fermi_background}. We take these uncertainties into account.
The model consists of two main contributions: the "atomic hydrogen" component produced by interactions of cosmic rays with interstellar matter and "inverse Compton" (IC) component produced by inverse Compton emission from cosmic ray electrons. Extrapolation of the atomic hydrogen component to the highest LAT energies is straightforward: the pion decay spectrum follows the cosmic ray spectrum and extrapolation has the form of a simple powerlaw with photon index $\sim 2.7$, the same as the slope of the cosmic ray spectrum. This component gives a sub-dominant contribution above 100~GeV. Extrapolation of the IC component depends on the unknown shape of the (average over interstellar medium) cosmic ray electron spectrum at the energies above TeV. We have checked that in the model of \citet{fermi_background} the spectrum of IC component is consistent with the spectrum of IC scattering of the local interstellar radiation field \citep{moskalenko06} by electrons with the spectrum $dN_e/dE\sim E^{-3}\exp\left(-E/1\mbox{ TeV}\right)$. This electron spectrum consistent with the cosmic ray electron spectrum observed on the Earth \citep{fermi_electrons,HESS_electrons}. The IC spectrum produced by such electron population is shown by the cyan dotted line in Fig. \ref{fig:Galactic}. The overall Galactic diffuse emission spectrum at high Galactic latitudes is then the sum of the atomic hydrogen and IC contributions, shown by the dashed black line in Fig. \ref{fig:Galactic}.
The high-energy cut-off of the local cosmic ray electron spectrum is most likely determined by the distance to the closest cosmic ray electron sources, e.g. to the closest pulsar wind nebulae \citep{aharonian04}, rather than by the intrinsic cut-off in the injection spectrum of electrons from the sources. This means that the local measurement of the high-energy cut-off of the cosmic electron spectrum does not provide a measurement of the high-energy cut-off in the injection spectrum of electrons. It is possible that Galactic cosmic electron sources inject electrons with energies much higher than $\sim 1$~TeV. This possibility is shown by the solid grey line in Fig. \ref{fig:Galactic} which shows the sum of the atomic hydrogen contribution with the IC emission from electrons without high-energy cut-off in the spectrum. The IC component still exhibits suppression at the energies $\sim 1$~TeV because of the Klein-Nishina effect.
\begin{figure}
\includegraphics[width=\linewidth]{fig2}
\caption{Extrapolation of the spectrum of the Galactic diffuse emission at high Galactic latitudes $|b|\ge 60^\circ$ to the 100~GeV energy range. Blue and cyan dotted lines below 100~GeV show the contributions from cosmic ray interactions with interstellar medium and inverse Compton scattering from cosmic electrons calculated by \citet{fermi_background}. Continuation of the cyan dotted line above 100~GeV is calculated assuming that inverse Compton emission is produced by electrons with a cut-off powerlaw spectrum with cut-off at 1~TeV. Grey dashed line is the sum of the cosmic ray and inverse Compton contributions. Solid grey line shows the overall diffuse emission spectrum in which the inverse Compton emission is produced by electron distribution without high-energy cut-off at 1~TeV. Red thick solid line shows the spectrum used for subtraction of Galactic component from the overall high Galactic latitude diffuse emission flux. }
\label{fig:Galactic}
\end{figure}
To take into account the above mentioned uncertainty of the IC component we adopt an approximation $dN_\gamma/dE\sim E^{-2.5}\exp\left(-E/2\mbox{ TeV}\right)$ for the high Galactic latitude diffuse emission spectrum at $E>100$~GeV. This approximation is shown by the red thick solid line in Fig. \ref{fig:Galactic}. This spectrum lies exactly in the middle between the two extreme possibilities: TeV-scale high-energy cut-off in the cosmic electron spectrum and no cut-off in the cosmic electron spectrum. One should take into account that the uncertainty of this approximation reaches $\simeq 50$\% at the highest energies. We take this uncertainty into account in the calculation of the EGB spectrum, by adding it as a systematic error. The two extreme possibilities for the behavior of electron spectrum above 1~TeV (exponential cut-off exactly at 1~TeV and no cut-off at all) provide a good estimate of the overall uncertainty of electron spectrum in the interstellar medium in this range. The uncertainty of the shape of electron spectrum dominates the uncertainty of the inverse Compton component of Galactic diffuse emission at high Galactic latitudes.
\subsection{Residual cosmic ray background contribution}
\label{sec:cr}
To estimate the residual cosmic ray background in the set of events selected for the analysis, we rely on the knowledge of residual
The residual cosmic ray background in the {\tt dataclean} event class of Pass 6 data is extensively discussed in \citet{fermi_background}. The residual cosmic ray background in the subset of Pass 7 {\tt superclean} events used in our analysis could be calculated from the known residual cosmic ray background in the Pass 6 {\tt dataclean} events via a straightforard comparison of statistics of events on- and off-point-sources in the two classes.
First, the residual cosmic ray fraction in the Pass 6 {\tt dataclean} events should be calculated from the known suppression factor of cosmic ray event at transition from the {\tt diffuse} event class to the {\tt dataclean} event class in Pass 6 (see \citet{fermi_background} for the detailed discussion of the suppression factor).
In each of the two event classes, the entire event set consists of a certain number of $\gamma$-ray\ events $N_{\gamma, i}$ and a certain number of residual cosmic ray events, $N_{CR, i}$ where $i$ stands for ${3+4}$ or $4$. Cleaning of the event set done to produce the {\tt dataclean} event set from {\tt diffuse} set results in rejection of a large fraction of the cosmic ray events, $N_{CR,4}=\alpha_{CR} N_{CR,3+4}$ with $\alpha_{CR}\ll 1$. However, it results also in rejection of a number of true $\gamma$-ray\ events, so that $N_{\gamma,4}=\alpha_\gamma N_{\gamma,3+4}$ with $\alpha_\gamma<1$.
The suppression factor $\alpha_{CR}$ is known as a function of energy from the Monte-Carlo simulations of cosmic ray and $\gamma$-ray\ induced events in the LAT detector by \citet{fermi_background}. The suppression factor $\alpha_\gamma$ could be found directly from the data set, by comparing statistics of events coming from the point sources in the {\tt diffuse} and {\tt dataclean} event classes (see section \ref{sec:ps} above). In the calculation of $\alpha_\gamma$ all the photons associated to $\sim 10^3$ point sources listed in the Fermi 2-year catalog \citep{fermi_catalog} could be used. This provides very large event statistics so that uncertainty of $\alpha_\gamma$ is negligible. Knowing the total numbers of events in the two event classes $N_{tot,i}$ one can resolve the system of equations
\begin{equation}
\left\{
\begin{array}{l}
N_{CR,4}/\alpha_{CR}+N_{\gamma,4}/\alpha_{\gamma}=N_{tot,3+4}\\
N_{CR,4}+N_{\gamma,4}=N_{tot,4}
\end{array}
\right.
\end{equation}
with respect to $N_{\gamma,4}, N_{CR,4}$ to find the residual cosmic ray background in each energy bin for the {\tt dataclean} event class.
The residual cosmic ray fraction in the sub-class of the Pass 7 {\tt superclean} events used in our analysis is then estimated in a similar way, once the residual cosmic ray fraction $\kappa_4$ in the point-source-subtracted set of the Pass 6 events, $N_{CR,\rm off,4}=\kappa_{4}N_{\rm off,4}$, is known.
Indeed, the transition from the Pass 6 {\tt dataclean} events to the Pass 7 events belonging to the event classes 65311 and 32543 leaves a fraction $\alpha_{\gamma,6\rightarrow 7}$ of $\gamma$-ray\ events (actually, $\alpha_{\gamma, 6\rightarrow 7}>1$ in a broad energy range around 10 GeV). It also suppresses or increases the residual cosmic ray background, so that the residual cosmic ray fraction in the off-source events changes from $\kappa_4$ to $\kappa_{7}$. The off-source events in the two classes are then the sum of the diffuse $\gamma$-ray\ emission photons and of the residual cosmic rays:
\begin{equation}
\left\{
\begin{array}{l}
\kappa_4 N_{\rm off,4}+N_{\gamma,\rm off, 4}=N_{\rm off,4}\\
\kappa_{7} N_{\rm off, 7}+\alpha_{\gamma, 6\rightarrow 7}N_{\gamma,\rm off, 4}=N_{\rm off,7}\\
\end{array}
\right.
\end{equation}
Knowing the statistics of the off-source events in the Pass 6 and Pass 7 events, $N_{\rm off,4}$ and $N_{\rm off,7}$, one could find the residual cosmic ray fraction in the Pass 7 data
\begin{equation}
\kappa_7=1-\alpha_{\gamma,6\rightarrow 7}(1-\kappa_6)\frac{N_{\rm off,4}}{N_{\rm off,7}}
\end{equation}
The resulting estimates of the level of residual cosmic ray background for the events selected in the Pass 7 data in energy bins between 3 and 100~GeV are shown by the grey data points in Fig. \ref{fig:spectrum}.
From Fig. \ref{fig:spectrum} one could see that the contribution of the residual cosmic rays to the signal at 100~GeV is likely to be small. However, extrapolation of the estimate of efficiency of rejection of the residual cosmic ray background much above 100~GeV is highly uncertain. It is possible that the efficiency of rejection of both the nuclear and electron/positron component of the cosmic ray flux drops because of the similarity of the cosmic ray and $e^+e^-$ pair tracks with large Lorentz factors. Inefficient rejection of the residual cosmic rays might lead to the contaminate the diffuse background signal and lead to a large over-estimation of the diffuse background flux. Because of this problem, we are able to only derive an upper limit on the EGB at the energies much above 100~GeV (for the energy band at 100~GeV we show a comparison between the 95\% confidence level upper limit and the measurement). A proper measurement of the EGB flux at the highest energies accessible to LAT would require extensive Monte-Carlo simulations taking into account detector response \citep{ackerman_texas}.
\subsection{Extragalactic $\gamma$-ray\ background spectrum}
\label{sec:spectrum}
EGB flux could be found by subtracting the point source, Galactic diffuse and residual cosmic ray contributions to the overall number of events at high Galactic latitudes in each energy bin. The spectrum of EGB obtained in this way is shown by the red thick data points in Fig. \ref{fig:spectrum}. The error in the measurement of the EGB flux at the energies below 100~GeV has contributions from the uncertainty of the level of the residual cosmic ray background as well as from the systematic uncertainties of the Instrument Response Functions (IRF) and of the Galactic diffuse background in the relevant sky region. An additional contribution to the error in the VHE band is given by the statistical error arising from the low signal statistics. Finally, one more uncertainty stems from the uncertainty of the shape of the cosmic ray electron spectrum in the TeV energy range, which propagates to the uncertainty of extrapolation of the inverse Compton component of the Galactic diffuse emission above 100 GeV.
\begin{figure}
\includegraphics[height=\columnwidth]{fig3}
\caption{Estimate of the flux of isotropic component of diffuse emission obtained by the direct photon counting method (red thick line, data points and upper limits). For comparison, the spectrum of isotropic component of diffuse sky emission, obtained using likelihood analysis at lower energies by \citet{fermi_background}, is shown as a red shaded region.
Pink upper limits above 100~GeV are from \citet{ackerman_texas}. Black data points show the total point-source-subtracted flux from the North and South Galactic pole regions at $|b|\ge 60^\circ$. Solid line errorbars show statistical error. Dashed line errorbar show the systematic error at the level of $\simeq 20$\% stemming from the uncertainty of the the Instrument Response Functions (IRF) of LAT (see {\tt http://fermi.gsfc.nasa.gov/ssc/data/analysis/LAT\_caveats.html}). Black horizontally shaded region shows the point source subtracted flux in the Galactic Pole regions found by \citet{fermi_background}. Grey data points and grey curve show the estimate of the residual cosmic ray background in the event set used in this analysis. The residual cosmic ray background level in the data set considered by \citet{fermi_background} is shown by the grey shaded region. Blue shading shows the Galactic diffuse emission in the North/South Galactic
Pole regions $|b|>60^\circ$ derived by \citet{fermi_background}. Blue line shows the Galactic diffuse background spectrum.}
\label{fig:spectrum}
\end{figure}
In the same figure we compare the measurement of EGB obtained from the direct photon counting at high Galactic latitudes with the results of the likelihood analysis of the all-sky data by \citet{fermi_background}. The two measurements agree well.
Pink arrows at the energies above 100~GeV show the upper limit on the VHE EGB derived by \citep{ackerman_texas} using the likelihood analysis of the all-sky data. These upper limits agree with the upper limits derived from the direct photon counting (shown by the red arrows in Fig. \ref{fig:spectrum}.
As a matter of fact, the level of Galactic diffuse emission at high Galactic latitudes turns out to be comparable to the level of EGB in the entire energy range $E>10$~GeV.\footnote{There is no a-priori reason why the two fluxes should be nearly equal. Thus, the equality of the two contributions poses a "fine-tuning" problem which requires further investigation. }
The Galactic diffuse emission contribution to the total flux at high Galactic latitude could not be negligibly small in the 100~GeV band. Taking this into account, it is not surprising that our estimate of the VHE EGB flux is somewhat lower than the total diffuse emission flux at high Galactic latitudes and is, respectively, lower than the upper limit derived by \citet{ackerman_texas}.
It is useful to note that extrapolating the EGB spectrum as a powerlaw spectrum to $E\ge 100$~GeV band would give the spectrum consistent with the data above 100~GeV. With the current LAT exposure, there is still no evidence for suppression of VHE EGB flux due to absorption on EBL. A larger exposure time is needed to verify the presence of the feature. As it is mentioned in the Introduction, suppression of the flux above 100~GeV due to the absorption of VHE $\gamma$-ray s on the EBL is expected if the EGB is accumulated over the cosmological distance scale. Detection of such suppression would be an important test of the origin of EGB.
\section{$E>30$~GeV Extragalactic $\gamma$-ray\ background from point sources}
As it is mentioned in the Introduction, different types of point and diffuse sources could contribute to the EGB in the VHE band. The main class of extragalactic point sources detected by {\it Fermi} is blazars, which are divided onto two sub-classed: BL Lac type objects and Flat Spectrum Radio Quasars (FSRQ). Over the first year of operation LAT has detected some $\sim 700$ such objects above 100~MeV energy \citep{fermi_agn}. BL Lacs and FSRQ have somewhat different spectral characteristics in the $\gamma$-ray\ band, with the spectra of BL Lacs being systematically harder than the spectra of FSRQ \citep{fermi_agn}. Hardness of the spectra of BL Lacs implies that they might produce significant contribution to the overall $\gamma$-ray\ flux in the VHE band. In fact, most of the extragalactic VHE $\gamma$-ray\ sources detected up to now by the ground based $\gamma$-ray\ telescopes sensitive above 100~GeV are BL Lacs\footnote{For the catalogs of extragalactic VHE $\gamma$-ray\ sources see e.g. {\tt http://tevcat.uchicago.edu} and {\tt http://www.isdc.unige.ch/vhe/index.html}}.
Fig. \ref{fig:sources} shows the breakdown of the point source contributions to the high Galactic latitude flux by the source type. One could clearly see that the dominant contribution is given by BL Lac objects which provide $\ge 90\%$ of the total point source flux above $30$~GeV. The cumulative spectrum of the other major blazar class, FSRQ has a high-energy cut-off at $\sim 10$~GeV so that FSRQ contribution to the point source flux is negligible in the VHE band. From Fig. \ref{fig:sources} one could see that the total point source flux calculated from the cumulative photon distribution around stacked point sources in the high Galactic latitude regions (see Section \ref{sec:ps}) is in a good agreement with the total point source flux calculated using the likelihood analysis by \citet{fermi_background}, shown by the green shaded region.
Above 50 GeV $ 90\%$ of source photons come from BL Lacs and $10\% $ from "Other" Fermi sources, which are dominated by not-identified sources with some contribution from nearby AGN's.
\begin{figure}
\includegraphics[height=\columnwidth]{fig4}
\caption{Cumulative $\gamma$-ray\ flux from different classes of resolved point sources. Source class is marked to the left from each curve. Green shaded area shows the point source flux at high Galactic latitudes found by \citet{fermi_background}. Red curve shows the EGB spectrum from Fig. \ref{fig:spectrum}, which includes only unresolved sources. }
\label{fig:sources}
\end{figure}
The EGB spectral shape above 30~GeV follows the cumulative point source spectrum. This observation leads to a conjecture that the VHE EGB is produced by already known type of VHE $\gamma$-ray\ point sources with fluxes below the sensitivity of LAT. Since the dominant source class in the VHE band is that of BL Lacs, a more precise conjecture is that the VHE EGB is produced by the unresolved BL Lacs.
\section{BL Lac contribution to VHE EGB}
In the unification schemes of AGN BL Lac objects are identified with the Fanaroff-Riley type I (FR I) radio galaxies with jets aligned with the line of sight \citep{urry95}. This implies that the cosmological evolution of the BL Lacs should follow that of the FR I radio galaxies. Recent studies of the cosmological evolution of FR I galaxies show that they experience "positive" cosmological evolution, which is usually described in terms of luminosity or comoving source density evolution as the increase of either average source luminosity or the average comoving source density with the redshift $z$, as $(1+z)^k,\ \ k>0$. Different recent studies find somewhat different values of $k$, depending on the analyzed radio galaxy samples and different assumptions about the evolution type (luminosity or density), with $k$ ranging in $1\lesssim k\lesssim 3$ \citep{sadler07,hodge09,smolcic09}. Since BL Lacs are just the FR I galaxies specially oriented with respect to the line of sight, their cosmological evolution follows the evolution of FR I galaxies, with the increasing source luminosity or spatial density with the redshift. This implies that significant flux should be produced by the sources at large redshifts, $z\sim 1$. It is possible that most of individual sources at large redshifts are too weak to be significantly detected by LAT, but collective emission from all the set of BL Lacs at high redshifts gives a significant contribution to the EGB.
\begin{figure}
\includegraphics[angle=-90,width=\columnwidth]{fig5}
\caption{Number of detected VHE photons as a function of redshift in the 6.25-12.5~GeV (blue dotted histogram) 25-50~GeV (green dashed histogram) and 100-200~GeV (red solid histogram) bands. Also shown are the distributions of photons with the redshift expected for different laws of BL Lac cosmological evolution. }
\label{fig:dNz}
\end{figure}
Dependence of the total flux of the BL Lac population on the redshift could be found from the following straightforward calculation.
Let us consider the total flux produced by sources at redshift $z$ in a redshift interval $\Delta z$. This redshift interval corresponds to the comoving distance interval $\Delta r=\Delta z/H(z)$ where $H(z)\sim\sqrt{\Omega_\Lambda+\Omega_m(1+z)^3}$ is the expansion rate of the Universe filled with matter and cosmological constant with today's densities $\Omega_m$ and $\Omega_\Lambda$.
As an example, we take the case of "pure luminosity" evolution with the average source luminosity increasing as $(1+z)^k$ and conserved comoving source density $n(z)=n_0=const$. The number of sources in a spherical layer of thickness $\Delta z$ is $\Delta N_s=4\pi n_0 r^2\Delta r$. Each source produces the flux in a given energy band $F\sim (1+z)^{k-\Gamma}/(4\pi r^2)$, where $\Gamma$ is the photon index, the factor $(1+z)^{1-\Gamma}$ describes the change in the number of photons in a given energy band due to the cosmological redshift of the photon energies. One power of $(1+z)$ is compensated by the time delay between subsequent photons.
The flux from the sources at large redshifts is affected by absorption of VHE photons on EBL. For example, at $z\simeq 1.5$ the absorption modifies source spectrum above the energy $E\simeq 50$ GeV, if one assumes the EBL evolution calculated by \citet{franceschini08}. Absorption on EBL leads to suppression of the flux by a factor $\exp\left(-\tau(E,z)\right)$ where
$\tau(E,z)$ is the optical depth with respect to the pair production.
The overall flux from the sources in the redshift interval $\Delta z$ is
\begin{equation}
\frac{\Delta F(E,z)}{\Delta z}=F\Delta N_s\sim \frac{(1+z)^{k-\Gamma}e^{-\tau(E,z)}}{\sqrt{\Omega_\Lambda+\Omega_m(1+z)^3}}
\label{dfdz}
\end{equation}
Fig. \ref{fig:dNz} shows the number of $\gamma$-ray s as a function of source redshift. For this we used BL Lacs with known redshifts from Veron\&Veron \citep{veron13} catalog complemented by BL Lacs detected by LAT, but not listed in the Veron\&Veron catalog. Only sources with $|b|>10^\circ$ were considered.
Here we plot photon distributions from Fermi BL Lacs in the three energy bands: $6.25-12.5$ GeV, $25-50$ GeV and $100-200$ GeV. One can see that at lower energies $E<50$ GeV a significant flux is produced by BL Lacs at large redshifts up to $z=1.5$. At the highest energies only contribution from nearby sources at $z<0.7$ is present. Two effects might explain the deficit of high-redshift sources at high energies. First, the flux at the highest energies is suppressed by absorption on EBL. Next, the photon statistics in the highest energy bin is low so that sources contributing to the flux in the 6.25-12.5~GeV bin produce less than one photon in the 100-200~GeV bin.
In the same figure we also show the expected dependence of the number of photons on the redshift expected in different evolution models, Eq.~(\ref{dfdz}). The models for cases $k=1,2,3$ are shown with magenta lines for $6.25-12.5$ GeV energy band. We normalize the models to the number of photons in the first redshift bin, in which we have the most complete knowledge of the BL Lac population.
From the comparison of the evolution models with the data one might get an impression that $(1+z)$ model is more consistent with the data than the models assuming faster evolution. However, the histogram on Fig. \ref{fig:dNz} does not take into account photons from BL Lacs with unknown redshifts. These BL Lacs produce about 30 \% of all cumulative BL Lac flux. This means that at least 30\% contribution to the overall flux (integrated over all redshifts) is missing in Fig. \ref{fig:dNz}.
The model with evolution $k=1$ predicts the total number of photons which is $\sim 3\sigma$ below the total number of photons in BL Lacs with known and unknown redshift together in the $6.25-12.5$ GeV energy band. In the energy band 3.125-6.25~GeV the under-prediction of the total number of photons from BL Lacs in the $k=1$ model is at $\ge 5\sigma$ level, which means that the model is efficiently ruled out. Thus, Fermi LAT observations of BL Lac objects indicate that BL Lacs have positive cosmological evolution with $k>1$.
In all other cases $k>1$, the discrepancy between the observed and expected number of photons from BL Lacs starts already at small redshifts, $z\ge 0.2$. The "missing BL Lac" $\gamma$-ray s could come either form BL Lacs with unknown redshifts or from {\it Fermi} sources which are not yet identified as BL Lacs or, finally, from BL Lacs with fluxes below the sensitivity of LAT.
\begin{figure}
\includegraphics[height=\columnwidth]{fig6}
\caption{VHE EGB produced by unresolved BL Lacs under different assumptions about the cosmological evolution of BL Lac population (the evolution law is marked to the left of each curve). Red data points show the EGB spectrum from Fig. \ref{fig:spectrum}.}
\label{fig:evolution}
\end{figure}
Although individual high redshift BL Lacs would not be detectable by LAT, cumulative flux of these BL Lacs could give significant contribution to the EGB. Fig. \ref{fig:evolution} shows the contributions from "missing BL Lacs" at high redshifts up to $z=1$ expected in four different models of cosmological evolution of BL Lac / FR I population. To calculate this contribution, we have normalized $\Delta F/\Delta z$ distribution on the measured flux of BL Lacs in the redshift bin $0<z<0.1$. For the two last bins,100-200~GeV and 200-400~GeV, the statistics of the BL Lac signal is too low to normalize $\Delta F/\Delta z$ by the flux in the first redshift bin. To estimate the normalization of $\Delta F/\Delta z$ in these energy bins we have assumed that the average BL Lac spectrum extends as a powerlaw with photon index $-2$ up to the 400~GeV energy range. The resulting statistics of the signal in the last 200-400 bin in Fig. \ref{fig:evolution} is low and is subject to large fluctuations.
From Fig. \ref{fig:evolution} one could see that the conjecture that EGB is produced by collective emission from distant BL Lacs is valid if BL Lacs experience positive luminosity or density evolution of the form $(1+z)^k$ with $k\simeq 3$. Slower or faster cosmological evolution with $k=2$ or $k=4$, respectively, would under- or over-produce the EGB. Since evolution $k = 4$ predicts number of photons more then
diffuse gamma-ray background (see Fig. \ref{fig:evolution}) it is also excluded. Thus, the LAT data impose a constraint on the cosmological evolution of the BL Lac/ FR I population
\begin{equation}
1< k < 4
\end{equation}
From Fig. \ref{fig:evolution} one could also see that if $k=3$, all the flux of EGB above 30~GeV could be explained by a cumulative emission from the BL Lac population. At the same time, if $k$ is significantly smaller than 3, as it is suggested by some recent studies of the evolution of the parent population of FR I galaxies \citep{smolcic09}, significant part of the VHE EGB flux should come from a yet unknown source population or have a truly diffuse nature.
$\gamma$-ray\ flux from high-redshift BL Lacs is modified in the VHE band by the effect of absorption on EBL. The model spectra shown in Fig. \ref{fig:evolution} take this effect into account. We use the model of \citet{franceschini08} to estimate the attenuation of the VHE $\gamma$-ray\ flux from redshifts up to $z=1.5$. Model curves shown in Fig. \ref{fig:evolution} assume that the average intrinsic spectrum of BL Lacs does not have a high-energy cut-off up to $\sim 400$~GeV. The observed suppression of the flux above 100~GeV is explained only by the effect of absorption on EBL.
\section{Discussion and Conclusions}
In this paper we have derived a measurement of EGB in the 10-400~GeV energy range from the analysis of {\it Fermi}/LAT data in the North and South Galactic Poles regions, $|b|>60^\circ$. Our approach was to count all the protons detected by LAT in this region and estimate the number of counts from the point sources, from the Galactic diffuse emission, form the residual cosmic ray background and from the EGB. Subtracting the source, Galactic diffuse and residual cosmic ray background counts from the total number of counts in the North and South Galactic Pole regions we derived the EGB spectrum shown in Fig. \ref{fig:spectrum}.
Comparing the spectrum of EGB in the $>30$~GeV energy band with the spectrum of extragalactic point sources in the same energy band (Fig. \ref{fig:sources}), we have noticed that the two spectra closely follow each other. Based on this observation, we have put forward a conjecture that the EGB above 30~GeV is explained by the unresolved BL Lacs, which give dominant contribution to the extragalactic point source flux in this energy band. We have demonstrated that this conjecture is consistent with the EGB measurement provided that BL Lacs follow positive cosmological evolution with the overall power of emission from the source population increasing as $(1+z)^3$ up to $z\sim 1$ (Fig. \ref{fig:evolution}).
Such cosmological evolution is roughly consistent with the measurements of cosmological evolution of FR I radio galaxies which are believed to be the parent population of BL Lac type objects and are also observed to have positive cosmological evolution of the form $(1+z)^k$ with an uncertain value of $k$ between 1 and 3 \citep{rigby08,smolcic09,sadler07}. At the same time, it is opposite to the negative cosmological evolution of the high-energy-peaked BL Lacs \citep{giommi99,giommi01}, which constitue a sub-class of the GeV-TeV $\gamma$-ray\ emitting BL Lacs considered in our analysis.
If the VHE EGB is indeed produced by distant BL Lacs at redshifts up to $z\sim 1$, LAT will not be able to resolve it into point sources. Indeed, the brightest BL Lac on the sky, Mrk 421 produced $\simeq 30$ photons above 100~GeV. If the positive cosmological evolution of BL Lac population is mostly due to the increase of the comoving source density rather than increase of the typical source luminosity, the brightest BL Lacs at redshift $z\sim 1$ produce $\sim 10^{-2}$ $\gamma$-ray s in LAT over some 2.5 years of exposure. This means that LAT would not collect sufficient photon statistics to detect distant BL Lacs individually. If the $(1+z)^3$ evolution is mostly due to the increase of the average source luminosity, BL Lacs at redshift 1 have an order-of-magnitude higher luminosity than local BL Lacs. However, even with higher luminosity, they produce on average 0.1 photon in LAT, so that they are still not individually detectable.
If the real value of $k$ is much below $k=3$, as indicated by a recent study by \citet{smolcic09}, emission from the unresolved BL Lacs would not explain the VHE EGB flux and there should be another source class or a mechanism of production of diffuse emission which would account for the EGB. Such a mechanism might, in fact, be indirectly related to the BL Lac population. Most of the power output from BL Lacs at the energies above 100~GeV is converted into the electromagnetic emission from $\gamma$-ray\ induced cascade in intergalactic medium. The intrinsic spectra of BL Lacs (and of FR I radio galaxies, such as M87 and Cen A \citep{m87,cena}) are known extend up to $\sim 10$~TeV. Typical energy of the cascade photons which are produced via inverse Compton scattering of CMB photons by the $e^+e^-$ pairs deposited in the intergalactic medium is $E_\gamma\simeq 100\left[E_{\gamma_0}/10\mbox{ TeV}\right]^2\mbox{ GeV}$ \citep{neronov09}. If the intrinsic source luminosities in the 1-10~TeV range are comparable to the luminosities in the 10-100~GeV, total flux of the cascade emission in the 10-100~GeV band is expected to be comparable to the point source flux, so that the cascade emission could give significant contribution to the EGB \citep{coppi97}. This is consistent with the observation that the VHE EGB level is comparable to the cumulative extragalactic point source flux in the 10-100~GeV band observed by LAT.
The only possibility to test the hypothesis of BL Lac origin of EGB would be to use deep observations with ground-based $\gamma$-ray\ telescopes. Ground based $\gamma$-ray\ telescopes, which are sensitive in the VHE energy band, have much larger collection area above several hundreds of GeV and, as a consequence, could detect much weaker sources, than LAT. The flux from Mrk 421-like BL Lacs at the redshift $z\simeq 1$ is $\sim 10^{-13}$~erg/cm$^2$s. If the cosmological evolution of BL Lacs is due to increase of the average source luminosity with redshift, brightest BL Lacs at redshift $z\sim 1$ might produce fluxes up to $10^{-12}$~erg/cm$^2$s at 100~GeV. At the energies around $E\lesssim 100$~GeV this flux is not strongly attenuated by the absorption on EBL.
The energy $E\simeq 100$~GeV is around or below the low energy threshold of the current generation Cherenkov telescopes, like HESS, MAGIC and VERITAS. However, next generation facilities, Cherenkov Telescope Array (CTA) \citep{cta} or 5@5 \citep{5at5} are expected to have an energy threshold significantly below 100~GeV. Their sensitivity could be sufficient to resolve the VHE EGB into point sources, at least in the case when the cosmological evolution is mostly luminosity, rather than source density evolution.
\section*{Acknowledgements}
We would like to thank I.Moskalenko, for the discussion of the issues related to the Galactic $\gamma$-ray\ background, and M.Ackermann for the clarification of the uncertainties of the residual cosmic ray background. The work of AN is supported by the Swiss National Science Foundation grant PP00P2\_123426.
|
\section{Introduction}
A language is \emph{prefix-free} (respectively, \emph{suffix-free}, \emph{factor-free}) if it does not contain any pair of words such that one is a proper prefix (respectively, suffix, factor) of the other. It is \emph{bifix-free} if it is both prefix- and suffix-free.
We refer to prefix-, suffix-, bifix-, and factor-free languages as \emph{free} languages.
Nontrivial prefix-, suffix-, bifix-, and factor-free languages are also known as prefix, suffix, bifix, and infix codes~\cite{BPR09,Shy01}, respectively and, have many applications in areas such as cryptography, data compression, and information processing.
The \emph{state complexity} of a regular language is the number of states in the minimal deterministic finite automaton (DFA) recognizing that language.
An equivalent notion is that of \emph{quotient complexity,} which is the number of left quotients of the language.
State complexity of regular operations has been studied quite extensively: for surveys of this topic and lists of references we refer the reader to~\cite{Brz09,Yu01}.
With regard to free regular languages, Han, Salomaa and Wood~\cite{HSW09} examined prefix-free regular languages, and Han and Salomaa~\cite{HS09} studied suffix-free regular languages.
Bifix- and factor-free regular languages were studied by Brzozowski, Jir\'askov\'a, Li, and Smith~\cite{BJLS11}.
The notion of quotient complexity can be derived from the Nerode right congruence~\cite{Ner58},
while the Myhill congruence~\cite{Myh57} leads to the syntactic semigroup of a language and to its \emph{syntactic complexity}, which is the cardinality of the syntactic semigroup.
It was pointed out in~\cite{BrYe11} that syntactic complexity can be very different for regular languages with the same quotient complexity.
Thus, for a fixed $n$, languages with quotient complexity $n$ may possibly be distinguished by their syntactic complexities.
In contrast to state complexity, syntactic complexity has not received much attention. In 1970 Maslov~\cite{Mas70} dealt with the problem of generators of the semigroup of all transformations in the setting of finite automata. In 2003--2004, Holzer and K\"onig~\cite{HoKo04}, and independently, Krawetz, Lawrence and Shallit~\cite{KLS03} studied the syntactic complexity of languages with unary and binary alphabets.
In 2010 Brzozowski and Ye~\cite{BrYe11} examined the syntactic complexity of ideal and closed regular languages, and in 2011 Brzozowski and Li~\cite{BL11} studied the syntactic complexity of star-free languages.
Here, we deal with the syntactic complexity of prefix-, suffix-, bifix-, and factor-free regular languages, and their complements.
Basic definitions and facts are stated in Sections~\ref{sec:trans} and~\ref{sec:complexity}. In Section~\ref{sec:pf} we obtain a tight upper bound on the syntactic complexity of prefix-free regular languages. In Sections~\ref{sec:sf}--\ref{sec:ff} we study the syntactic complexity of suffix-, bifix-, and factor-free regular languages, respectively. We state conjectures about tight upper bounds for these classes, and exhibit languages in these classes that have large syntactic complexities. In Section~\ref{sec:rev} we show that the upper bounds on the quotient complexity of reversal of prefix-, suffix-, bifix-, and factor-free regular languages can be met by our languages with largest syntactic complexities. Section~\ref{sec:cl} concludes the paper.
\section{Transformations}\label{sec:trans}
A {\em transformation} of a set $Q$ is a mapping of $Q$ into itself. In this paper we consider only transformations of finite sets, and we assume without loss of generality that $Q = \{1,2,\ldots, n\}$. Let $t$ be a transformation of $Q$. If $i \in Q$, then $it$ is the {\it image} of $i$ under $t$. If $X$ is a subset of $Q$, then $Xt = \{it \mid i \in X\}$, and the {\em restriction} of $t$ to $X$, denoted by $t|_X$, is a mapping from $X$ to $Xt$ such that $it|_X = it$ for all $i \in X$. The {\em composition} of two transformations $t_1$ and $t_2$ of $Q$ is a transformation $t_1 \circ t_2$ such that $i (t_1 \circ t_2) = (i t_1) t_2$ for all $i \in Q$. We usually drop the composition operator ``$\circ$'' and write $t_1t_2$ for short.
An arbitrary transformation can be written in the form
\begin{equation*}\label{eq:transmatrix}
t=\left( \begin{array}{ccccc}
1 & 2 & \cdots & n-1 & n \\
i_1 & i_2 & \cdots & i_{n-1} & i_n
\end{array} \right ),
\end{equation*}
where $i_k = kt$, $1\le k\le n$, and $i_k\in Q$. The {\em domain} $\mathop{\mbox{dom}}(t)$ of $t$ is $Q.$
The {\em range} $\mathop{\mbox{rng}}(t)$ of $Q$ under $t$ is the set
$\mathop{\mbox{rng}}(t) = Q t.$ We also use the notation $t = [i_1,i_2,\ldots,i_n]$ for the transformation $t$ above.
A \emph{permutation} of $Q$ is a mapping of $Q$ \emph{onto} itself. In other words, a permutation $\pi$ of $Q$ is a transformation where $\mathop{\mbox{rng}}(\pi) = Q$.
The \emph{identity} transformation maps each element to itself, that is, $it=i$ for $i=1,\ldots,n$.
A transformation $t$ contains a \emph{cycle} of length $k$ if there exist pairwise different elements $i_1,\ldots,i_k$ such that
$i_1t=i_2, i_2t=i_3,\ldots, i_{k-1}t=i_k$, and $i_kt=i_1$.
A cycle is denoted by $(i_1,i_2,\ldots,i_k)$.
For $i<j$, a \emph{transposition} is the cycle $(i,j)$, and $(i,i)$ is the identity.
A \emph{singular} transformation, denoted by $i\choose j$, has $it=j$ and $ht=h$ for all $h\neq i$, and $i \choose i$ is the identity.
A~\emph{constant} transformation, denoted by $Q \choose j$, has $it=j$ for all $i$.
The set of all transformations of a set $Q$, denoted by ${\mathcal T}_Q$, is a finite monoid. The set of all permutations of $Q$ is a group, denoted by $\mathfrak{S}_Q$ and called the \emph{symmetric group} of degree $n$.
It was shown in~\cite{Hoy1895,Pic38} that two generators are sufficient to generate the symmetric group of degree $n$.
In 1935 Piccard~\cite{Pic35} proved that three transformations of $Q$ are sufficient to generate the monoid ${\mathcal T}_Q$. In the same year, Eilenberg showed that fewer than three generators are not possible, as reported by Sierpi\'nski~\cite{Sie35}. We refer the reader to the book of Ganyushkin and Mazorchuk~\cite{GaMa09} for a detailed discussion of finite transformation semigroups. The following are well-known facts about generators of $\mathfrak{S}_Q$ and ${\mathcal T}_Q$:
\begin{theorem}[Permutations,~\cite{Hoy1895,Pic38}]
\label{thm:piccard}
The symmetric group $\mathfrak{S}_Q$ of size $n!$ can be generated by any cyclic
permutation of $n$ elements together with any transposition. In particular, $\mathfrak{S}_Q$ can be generated by
$c=(1,2,\ldots, n)$~and~$t~=~(1,2)$.
\end{theorem}
\begin{theorem}[Transformations,~\cite{Pic35}]
\label{thm:salomaa}
The complete transformation monoid ${\mathcal T}_Q$ of size $n^n$ can be generated by any cyclic
permutation of $n$ elements together with a transposition and a ``returning'' transformation $r={n \choose 1}$. In particular, ${\mathcal T}_Q$ can be generated by $c=(1,2,\ldots, n)$, $t=(1,2)$ and $r={n \choose 1}$.
\end{theorem}
\section{Quotient Complexity and Syntactic Complexity}\label{sec:complexity}
If $\Sigma$ is a non-empty finite alphabet, then $\Sigma^*$ is the free monoid generated by $\Sigma$, and $\Sigma^+$ is the free semigroup generated by $\Sigma$. A \emph{word} is any element of $\Sigma^*$, and the empty word is $\varepsilon$. The length of a word $w\in \Sigma^*$ is $|w|$. A \emph{language} over $\Sigma$ is any subset of $\Sigma^*$.
If $w=uxv$ for some $u,x,v\in\Sigma^*$, then $u$ is a {\em prefix\/} of $w$, $v$ is a {\em suffix\/} of $w$, and $x$ is a {\em factor\/} of $w$. Both $u$ and $v$ are also factors of $w$.
A~{\em proper} prefix (suffix, factor) of $w$ is a prefix (suffix, factor) of $w$ other than~$w$.
The \emph{left quotient}, or simply \emph{quotient,} of a language $L$ by a word $w$ is the language $L_w=\{x\in \Sigma^*\mid wx\in L \}$.
For any $L\subseteq \Sigma^*$, the \emph{Nerode right congruence}~\cite{Ner58} ${\hspace{.1cm}{\sim_L} \hspace{.1cm}}$ of $L$ is defined as follows:
\begin{equation*}
x {\hspace{.1cm}{\sim_L} \hspace{.1cm}} y \mbox{ if and only if } xv\in L \Leftrightarrow yv\in L, \mbox { for all } v\in\Sigma^*.
\end{equation*}
Clearly, $L_x=L_y$ if and only if $x{\hspace{.1cm}{\sim_L} \hspace{.1cm}} y$.
Thus each equivalence class of this right congruence corresponds to a distinct quotient of $L$.
The \emph{Myhill congruence}~\cite{Myh57} ${\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ of $L$ is defined as follows:
\begin{equation*}
x {\hspace{.1cm}{\approx_L} \hspace{.1cm}} y \mbox{ if and only if } uxv\in L \Leftrightarrow uyv\in L\mbox { for all } u,v\in\Sigma^*.
\end{equation*}
This congruence is also known as the \emph{syntactic congruence} of $L$.
The quotient set $\Sigma^+/ {\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ of equivalence classes of the relation ${\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ is a semigroup called the \emph{syntactic semigroup} of $L$, and
$\Sigma^*/ {\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ is the \emph{syntactic monoid} of~$L$.
The \emph{syntactic complexity} $\sigma(L)$ of $L$ is the cardinality of its syntactic semigroup.
The \emph{monoid complexity} $\mu(L)$ of $L$ is the cardinality of its syntactic monoid.
If the equivalence class containing $\varepsilon$ is a singleton in the syntactic monoid, then $\sigma(L)=\mu(L)-1$; otherwise, $\sigma(L)=\mu(L)$.
A~\emph{deterministic finite automaton} (DFA) is a quintuple ${\mathcal A}=(Q, \Sigma, \delta, q_1,F)$, where
$Q$ is a finite, non-empty set of \emph{states}, $\Sigma$ is a finite non-empty \emph{alphabet}, $\delta:Q\times \Sigma\to Q$ is the \emph{transition function}, $q_1\in Q$ is the \emph{initial state}, and $F\subseteq Q$ is the set of \emph{accepting states}. We extend $\delta$ to $Q \times \Sigma^*$ in the usual way.
The DFA ${\mathcal A}$ accepts a word $w \in \Sigma^*$ if ${\delta}(q_1,w)\in F$.
The set of all words {\it accepted} by ${\mathcal A}$ is $L({\mathcal A})$.
By the \emph{language of a state} $q$ of ${\mathcal A}$
we mean the language accepted
by the DFA $(Q,\Sigma,\delta,q,F)$.
A state is \emph{empty} if its language is empty.
Let $L$ be a regular language.
The \emph{quotient DFA} of $L$ is
${\mathcal A}=(Q, \Sigma, \delta, q_1,F)$, where $Q=\{L_w\mid w\in\Sigma^*\}$, $\delta(L_w,a)=L_{wa}$,
$q_1=L_\varepsilon=L$, $F=\{L_w \mid \varepsilon \in L_w\}$.
The number $\kappa(L)$ of distinct quotients of $L$ is the \emph{quotient complexity} of $L$.
The quotient DFA of $L$ is the minimal DFA accepting $L$, and so quotient complexity is the same as state complexity, but there are advantages to using quotients~\cite{Brz09}.
In terms of automata, each equivalence class $[w]_{{\hspace{.1cm}{\sim_L} \hspace{.1cm}}}$ of ${\hspace{.1cm}{\sim_L} \hspace{.1cm}}$ is the set of all words $w$ that take the automaton to the same state from the initial state, and each equivalence class $[w]_{\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ of ${\hspace{.1cm}{\approx_L} \hspace{.1cm}}$ is the set of all words that perform the same transformation on the set of states~\cite{McNP71}.
In terms of quotients, $[w]_{{\hspace{.1cm}{\sim_L} \hspace{.1cm}}}$ is the set of words $w$ that can be followed by the same quotient $L_w$.
Let ${\mathcal A} = (Q, \Sigma, \delta, q_1, F)$ be a DFA. For each word $w \in \Sigma^*$, the transition function for $w$ defines a transformation $t_w$ of $Q$ by the word $w$: for all $i \in Q$,
$it_w \stackrel{\rm def}{=} \delta(i, w).$
The set $T_{{\mathcal A}}$ of all such transformations by non-empty words forms a subsemigroup of ${\mathcal T}_Q$, called the \emph{transition semigroup} of ${\mathcal A}$~\cite{Pin97}.
Conversely, we can use a set $\{t_a \mid a \in \Sigma\}$ of transformations to define $\delta$, and so the DFA ${\mathcal A}$. When the context is clear we simply write $a = t$, where $t$ is a transformation of $Q$, to mean that the transformation performed by $a \in \Sigma$ is~$t$.
If ${\mathcal A}$ is the quotient DFA of $L$, then $T_{{\mathcal A}}$ is isomorphic to the syntactic semigroup $T_L$ of $L$~\cite{McNP71}, and we represent elements of $T_L$ by transformations in~$T_{{\mathcal A}}$.
We attempt to obtain tight upper bounds on the syntactic complexity $\sigma(L) = |T_L|$ of $L$ as a function of the quotient complexity $\kappa(L)$ of $L$.
First we consider the syntactic complexity of regular languages over a unary alphabet, where the concepts prefix-, suffix-, bifix-, and factor-free, coincide. So we may consider only unary prefix-free regular languages $L$ with quotient complexity $\kappa(L) = n$. When $n = 1$, the only prefix-free language is $L = \emptyset$ with $\sigma(L) = 1$. For $n \ge 2$, a prefix-free language $L$ must be a singleton, $L = \{a^{n-2}\}$. The syntactic semigroup $T_L$ of $L$ consists of $n-1$ transformations $t_w$ by words $w = a^i$, where $1 \le i \le n-1$. Thus we have
\begin{proposition}[Unary Free Regular Languages]
If $L$ is a unary free regular language with $\kappa(L) = n \ge 2$, then $\sigma(L) = n-1$.
\end{proposition}
The tight upper bound for regular unary languages~\cite{HoKo04} is $n$.
We assume that $|\Sigma| \ge 2$ in the following sections.
Since the syntactic semigroup of a language is the same as that of its complement, we deal only with prefix-, suffix-, bifix-, and factor-free languages. All the syntactic complexity results, however, apply also to the complements of these languages.
\goodbreak
\section{Prefix-Free Regular Languages}\label{sec:pf}
To simplify notation we write $\varepsilon$ for the language $\{\varepsilon\}$. Recall that a regular language $L$ is prefix-free if and only it has exactly one accepting quotient, and that quotient is $\varepsilon$~\cite{HSW09}.
\begin{theorem}[Prefix-Free Regular Languages]
\label{thm:prefix-free}
If $L$ is regular and prefix-free with $\kappa(L)=n\ge 2$, then $\sigma(L)\le n^{n-2}$. Moreover, this bound is tight
for $n=2$ if $|\Sigma|\ge 1$, for $n=3$ if $|\Sigma|\ge 2$, for $n=4$ if $|\Sigma|\ge 4$, and for $n\ge 5$
if $|\Sigma|\ge n+1$.
\end{theorem}
\begin{proof}
If $L$ is prefix-free, the only accepting quotient of $L$ is $\varepsilon$. Thus $L$ also has the empty quotient, since $\varepsilon_a = \emptyset$ for $a \in \Sigma$.
Let ${\mathcal A} = (Q, \Sigma, \delta, 1, \{n-1\})$ be the quotient DFA of $L$, where, without loss of generality, $n-1 \in Q$ is the only accepting state, and $n \in Q$ is the empty state. For any transformation $t \in T_L$, $(n-1) t =n t = n$. Thus we have $\sigma(L)\le n^{n-2}$.
The only prefix-free regular language for $n=1$ is $L=\emptyset$ with $\sigma(L)=1$; here the bound $n^{n-2}$ does not apply.
For $n=2$ and $\Sigma=\{a\}$, the language $L=\varepsilon$ meets the bound.
For $n=3$ and $\Sigma=\{a,b\}$, $L=b^*a$ meets the bound.
For $n\ge 4$, let ${\mathcal A}_n=(\{1, 2,\ldots, n\}, \{a,b,c,d_1,d_2,\ldots,d_{n-2}\},\delta,1,\{n-1\})$, where $a={{n-1} \choose n} (1,2,\ldots,n-2)$, $b={{n-1} \choose n} (1,2)$, $c={{n-1} \choose n} {n-2\choose 1}$, and $d_i={{n-1} \choose n} {i\choose n-1}$ for $i=1,2,\ldots,n-2$.
DFA ${\mathcal A}_6$ is shown in Fig.~\ref{fig:PrFree},
where $\Gamma = \{d_1,d_2,\ldots,d_{n-2}\}$.
For $n=4$, input $a$ coincides with $b$; hence only $4$ inputs are needed.
\begin{figure}[hbt]
\begin{center}
\input{PrFree.eepic}
\end{center}
\caption{Quotient DFA ${\mathcal A}_6$ of prefix-free regular language with 1,296 transformations.}
\label{fig:PrFree}
\end{figure}
Any transformation $t \in T_L$ has the form
$$t=\left( \begin{array}{cccccc}
1 & 2 & \cdots & n-2 & n-1 & n \\
i_1 & i_2 & \cdots & i_{n-2} & n & n
\end{array} \right ),
$$
where $i_k\in\{1,2,\ldots,n\}$ for $1\le k\le n-2$.
There are three cases:
\begin{enumerate}
\item If $i_k\le n-2$ for all $k$, $1\le k\le n-2$, then by Theorem~\ref{thm:salomaa}, ${\mathcal A}_n$ can do $t$.\\
\item If $i_k\le n-1$ for all $k$, $1\le k\le n-2$, and there exists some $h$ such that $i_h= n-1$, then there exists some $j$, $1\le j\le n-2$ such that $i_k\ne j$ for all $k$, $1\le k\le n-2$.
For all $1\le k\le n-2$, define $i'_k$ as follows: $i'_k = j$ if $i_k=n-1$, and $i'_k = i_k$ if $i_k\ne n-1$.
Let
$$
s=\left( \begin{array}{cccccc}
1 & 2 & \cdots & n-2 & n-1 & n \\
i'_1 & i'_2 & \cdots & i'_{n-2} & n & n
\end{array} \right ).$$
By Case 1 above, ${\mathcal A}_n$ can do $s$.
Since $t=sd_j$, ${\mathcal A}_n$ can do $t$ as well.\\
\item Otherwise, there exists some $h$ such that $i_h= n$. Then there exists some $j$, $1\le j\le n-2$, such that $i_k\ne j$ for all $k$, $1\le k\le n-2$.
For all $1\le k\le n-2$, define $i'_k$ as follows: $i'_k = n-1$ if $i_k=n$, $i'_k = j$ if $i_k=n-1$, and $i'_k = i_k$ otherwise.
Let $s$ be as above but with new $i'_k$.
By Case 2 above, ${\mathcal A}_n$ can do $s$.
Since $t=sd_j$, ${\mathcal A}_n$ can do $t$ as well.
\end{enumerate}
Therefore, the syntactic complexity of ${\mathcal A}_n$ meets the desired bound. \qed
\end{proof}
We conjecture that the alphabet sizes cannot be reduced. As shown in Table~\ref{tab:Summary1},
on p.~\pageref{table1},
we have verified this conjecture for $n \le 5$ by enumerating all prefix-free regular languages with $n\le 5$ using \emph{GAP}~\cite{GAP}.
\medskip
\section{Suffix-Free Regular Languages}\label{sec:sf}
For any regular language $L$, a quotient $L_w$ is \emph{uniquely reachable}~\cite{Brz09} if $L_w=L_x$ implies that $w=x$.
It is known from~\cite{HS09} that, if $L$ is a suffix-free regular language, then $L=L_\varepsilon$ is uniquely reachable by $\varepsilon$, and $L$ has the empty quotient.
Without loss of generality, we assume that $1$ is the initial state, and $n$ is the empty state.
We will show that the cardinality of $\Bsf(n)$, defined below, is an upper bound ($\mathbf{B}$ for ``bound'') on the syntactic complexity of suffix-free regular languages with quotient complexity $n$. Let
$$\Bsf(n) = \{ t \in {\mathcal T}_{Q} \mid 1 \not\in \mathop{\mbox{rng}}(t), \; nt = n, \txt{and for all} j\ge 1,
\hspace{2.5cm}$$
$$1t^j = n \txt{or} 1t^j \neq it^j ~~\forall i, 1 < i < n\}.$$
\begin{proposition}
\label{prop:sf}
If $L$ is a regular language with quotient DFA ${\mathcal A}_n = (Q, \Sigma, \delta, 1, F)$ and syntactic semigroup $T_L$, then the following hold:
\begin{enumerate}
\item If $L$ is suffix-free, then $T_L$ is a subset of $\Bsf(n)$.
\item If $L$ has the empty quotient, only one accepting quotient, and $T_L \subseteq \Bsf(n)$, then $L$ is suffix-free.
\end{enumerate}
\end{proposition}
\begin{proof}
1. Let $L$ be suffix-free, and let ${\mathcal A}_n$ be its quotient DFA.
Consider an arbitrary $t \in T_L$. Since the quotient $L$ is uniquely reachable, $it \neq 1$ for all $i \in Q$. Since the quotient corresponding to state $n$ is empty, $nt = n$.
Since $L$ is suffix-free, for any two quotients $L_w$ and $L_{uw}$, where $u,v,w \in \Sigma^+$, $w = v^j$ for some $j \ge 1$, and $L_w \neq \emptyset$, we must have $L_w \cap L_{uw} = \emptyset$, and so $L_w \neq L_{uw}$.
This means that, for any $t \in T_L$ and $j \ge 1$, if $1t^j \neq n$, then $1t^j \neq it^j$ for all $i$, $1 < i < n$. So $t \in \Bsf(n)$, and $T_L \subseteq \Bsf(n)$.
2. Assume that $T_L \subseteq \Bsf(n)$, and let $f$ be the only accepting state. If $L$ is not suffix-free, then there exist non-empty words $u$ and $v$ such that $v, uv \in L$. Let $t_u$ and $t_v$ be the transformations by $u$ and $v$, and let $i = 1t_u$; then $i \neq 1$.
Assume without loss the generality that $n$ is the empty state. Then $f\neq n$, and we have $1t_v = f = 1t_{uv} = 1t_ut_v = it_v$, which contradicts the fact that $t_v \in \Bsf(n)$. Therefore $L$ is suffix-free. \qed
\end{proof}
Let $\bsf(n) = |\Bsf(n)|$. We now prove that $\bsf(n)$ is an upper bound on the syntactic complexity of suffix-free regular languages.
With each transformation $t$ of $Q$, we associate a directed graph $G_t$, where $Q$ is the set of nodes, and $(i,j) \in Q \times Q$ is a directed edge from $i$ to $j$ if $it = j$. We call such a graph $G_t$ the {\em transition graph} of $t$. For each node $i$, there is exactly one edge leaving $i$ in $G_t$. Consider the infinite sequence $i,it,it^2,\ldots$ for any $i \in Q$. Since $Q$ is finite, there exists least $j \ge 0$ such that $it^{j+1} = it^{j'}$ for some $j' \le j$. Then the finite sequence ${\mathfrak{s}}_t(i) = i,it,\ldots,it^j$ contains all the distinct elements of the above infinite sequence, and it induces a directed path $P_t(i)$ from $i$ to $it^j$ in $G_t$. In particular, if $n \in {\mathfrak{s}}_t(1)$, and $nt = n$, then we call ${\mathfrak{s}}_t(1)$ the {\em principal sequence} of $t$, and $P_t(1)$, the {\em principal path} of $G_t$.
\begin{proposition}\label{prop:ppsf}
There exists a principal sequence for every transformation~$t$ in~$\Bsf(n)$.
\end{proposition}
\begin{proof}
Suppose $t \in \Bsf(n)$ and ${\mathfrak{s}}_t(1) = 1,1t,\ldots,1t^j$.
If $t$ does not have a principal sequence,
then $n \not\in {\mathfrak{s}}_t(1)$, and $1t^{j+1} = 1t^{j'} \neq n$ for some $j' \le j$.
Let $i=1t^{j+1-j'}$; then $i \neq 1$ and $1t^{j'} =it^{j'}$, violating the last property of $\Bsf(n)$.
Therefore there is a principal sequence for every $t \in \Bsf(n)$. \qed
\end{proof}
Fix a transformation $t \in \Bsf(n)$. Let $i \in Q$ be such that $i \not\in {\mathfrak{s}}_t(1)$. If the sequence ${\mathfrak{s}}_t(i)$ does not contain any element of the principal sequence ${\mathfrak{s}}_t(1)$ other than $n$, then we say that ${\mathfrak{s}}_t(i)$ has {\em no principal connection}. Otherwise, there exists least $j \ge 1$ such that $1t^j \neq n$ and $1t^j = it^{j'} \in {\mathfrak{s}}_t(i)$ for some $j' \ge 1$, and we say that ${\mathfrak{s}}_t(i)$ has a {\em principal connection} at $1t^j$. If $j' < j$, the principal connection is {\em short}; otherwise, it is {\em long}.
\begin{lemma}\label{lem:shortpath}
For all $t \in \Bsf(n)$ and $i \not\in {\mathfrak{s}}_t(1)$, the sequence ${\mathfrak{s}}_t(i)$ has no long principal connection.
\end{lemma}
\begin{proof}
Let $t$ be any transformation in $\Bsf(n)$. Suppose for some $i \not\in {\mathfrak{s}}_t(1)$, the sequence ${\mathfrak{s}}_t(i)$ has a long principal connection at $1t^j = it^{j'} \neq n$, where $j < j'$. Hence $it^{j'-j} \neq n$, and $1t^j = (it^{j'-j})t^j$, which is a contradiction. Therefore, for all $i \not\in {\mathfrak{s}}_t(1)$, ${\mathfrak{s}}_t(i)$ has no long principal connection. \qed
\end{proof}
To calculate the cardinality of $\Bsf(n)$, we need the following observation.
\begin{lemma}\label{lem:ptree}
For all $t \in \Bsf(n)$ and $i \not\in {\mathfrak{s}}_t(1)$, if ${\mathfrak{s}}_t(i)$ has a principal connection, then there is no cycle incident to the path $P_t(i)$ in the transition graph $G_t$.
\end{lemma}
\begin{proof}
This observation can be derived from Theorem 1.2.9 of~\cite{GaMa09}. However, our proof is shorter.
Pick any $i \not\in {\mathfrak{s}}_t(1)$ such that ${\mathfrak{s}}_t(i)$ has a principal connection at $1t^j = it^{j'}$ for some $i,j$ and $j'$. Then the sequence ${\mathfrak{s}}_t(i)$ contains $n$, and the path $P_t(i)$ does not contain any cycle. Suppose $C$ is a cycle which includes node
$x=it^k \in P_t(i)$.
Since there is only one outgoing edge for each node in $G_t$, the cycle $C$ must be oriented and must contain a node $x'\not\in P_t(i)$ such that $(x',x)$ is an edge in $C$.
Then the next node in the cycle must be $it^{k+1}$ since there is only one outgoing edge from $x$. But then $x'$ can never be reached from $P_t(i)$, and so no such cycle can exist. \qed
\end{proof}
By Lemma~\ref{lem:ptree}, for any $1t^j \in {\mathfrak{s}}_t(1)$, where $j \ge 1$, the union of directed paths from various nodes $i$ to $1t^j$, if $i \not\in {\mathfrak{s}}_t(1)$ and ${\mathfrak{s}}_t(i)$ has a principal connection at $1t^j$, forms a labeled tree $T_t(j)$ rooted at $it^j$. Suppose there are $r_j+1$ nodes in $T_t(j)$ for each $j$, and suppose there are $r$ elements of $Q$ that are not in the principal sequence ${\mathfrak{s}}_t(1)$ nor in any tree $T_t(j)$, for some $r_j,r \ge 0$. Note that, $it^j$ is the only node in $T_t(j)$ that is also in the principal sequence ${\mathfrak{s}}_t(1)$. Each tree $T_t(j)$ has height at most $j-1$; otherwise, some $i \in T_t(j)$ has a long principal connection. In particular, tree $T_t(1)$ has height 1; so it is trivial with only one node $1t$. Then $r_1 = 0$, and we need only consider trees $T_t(j)$ for $j \ge 2$. Let $S_m(h)$ be the number of labeled rooted trees with $m$ nodes and height at most $h$. This number can be found in the paper of Riordan~\cite{Rio60}; the calculation is somewhat complex, and we refer the reader to~\cite{Rio60} for details. For convenience, we include the values of $S_m(h)$ for small values of $m$ and $h$ in Table~\ref{tab:Smh}, where the row number is $h$ and the column number is $m$.
\begin{table}[ht]
\caption{The number $S_m(h)$ of labeled rooted trees with $m$ nodes and height at most~$h$.}
\label{tab:Smh}
\begin{center}
$
\begin{array}{|c||c|c|c|c|c|c|c|}
\hline
h/m & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\
\hline \hline
0 & 1 & 0 & 0 & 0 & 0 & 0 & 0 \\
\hline
1 & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\
\hline
2 & 1 & 2 & 9 & 40 & 205 & 1176 & 7399 \\
\hline
3 & 1 & 2 & 9 & 64 & 505 & 4536 & 46249\\
\hline
4 & 1 & 2 & 9 & 64 & 625 & 7056 & 89929 \\
\hline
5 & 1 & 2 & 9 & 64 & 625 & 7776 & 112609 \\
\hline
6 & 1 & 2 & 9 & 64 & 625 & 7776 & 117649 \\
\hline
\end{array}
$
\end{center}
\label{table0}
\end{table}
Since each of the $m$ nodes can be the root, there are $S'_m(h) = \frac{S_m(h)}{m}$ labeled trees rooted at a fixed node and having $m$ nodes and height at most $h$. The following is an example of trees $T_t(j)$ in transformations $t \in \Bsf(n)$.
\begin{example}\label{ex:ptree}
Let $n = 15$. Consider any transformation $t \in \Bsf(15)$ with principal sequence ${\mathfrak{s}}_t(1) = 1,2,3,4,5,15$. There are $9$ elements of $Q$ that are not in ${\mathfrak{s}}_t(1)$, and some of them are in the trees $T_t(j)$ for $2 \le j \le 4$. Consider the cases where $r_2 = 2$, $r_3 = 3$, $r_4 = 1$, and $r = 3$. Fig.~\ref{fig:ptree} shows one such transformation $t$.
\begin{figure}[hbt]
\begin{center}
\input{ptree.eepic}
\end{center}
\caption{Transition graph of some $t \in \Bsf(15)$ with principal sequence $1,2,3,4,5,15$.}
\label{fig:ptree}
\end{figure}
For $j = 2$, the tree $T_t(2)$ has height at most $1$, and there are $S'_{r_2+1}(1) = \frac{S_{r_2+1}(1)}{r_2+1} = \frac{3}{3} = 1$ possible $T_t(2)$. For $j=3$, there are $S'_{r_3+1}(2) = \frac{S_{r_3+1}(2)}{r_3+1} = 10$ possible $T_t(3)$, which are of one of the three types shown in Fig.~\ref{fig:Tj}. Among the 10 possible $T_t(3)$, one is of type (a), three are of type (b), and six are of type~(c). For $j = 4$, there are $S'_{r_4+1}(3) = \frac{S_{r_4+1}(3)}{r_4+1} = 1$ possible $T_t(4)$.
\begin{figure}[hbt]
\begin{center}
\input{Tj.eepic}
\end{center}
\caption{Three types of trees of the form $T_t(3)$, where $\{i_1,i_2,i_3\} = \{8,9,10\}$.}
\label{fig:Tj}
\end{figure}
\end{example}
Let $C^n_k$ be the binomial coefficient, and let $C^n_{k_1,\ldots,k_m}$ be the multinomial coefficient. Then we have
\begin{lemma}\label{lem:Gn}
For $n \ge 3$, we have
\begin{equation}\label{eq:g}
\bsf(n) =
\sum_{k=0}^{n-2} C^{n-2}_k k!
\sum_{\begin{subarray}{c}
r_2+\cdots+r_k+r\\
= n-k-2
\end{subarray}}
C^{n-k-2}_{r_2,\ldots,r_k,r} (r+1)^r \prod_{j=2}^k S'_{r_j+1}(j-1) .
\end{equation}
\end{lemma}
\begin{proof}
Let $t$ be any transformation in $\Bsf(n)$. Suppose ${\mathfrak{s}}_t(1) = 1,1t,\ldots,1t^k,n$ for some $k$, $0 \le k \le n-2$. There are $C^{n-2}_k k!$ different principal sequences ${\mathfrak{s}}_t(1)$. Now, fix ${\mathfrak{s}}_t(1)$. Suppose $n-k-2 = r_2 + \cdots + r_k + r$, where, for $2 \le j \le k$, tree $T_t(j)$ contains $r_j+1$ nodes, for some $r_j \ge 0$. There are $C^{n-k-2}_{r_2,\ldots,r_k,r}$ different tuples $(r_2,\ldots,r_k,r)$. Each tree $T_t(j)$ has height at most $j-1$, and it is rooted at $1t^j$. There are $S'_{r_j+1}(j-1) = \frac{S_{r_j+1}(j-1)}{r_j + 1}$ different trees $T_t(j)$. Let $E$ be the set of the remaining $r$ elements $x$ of $Q$ that are not in any tree $T_t(j)$ nor in the principal sequence ${\mathfrak{s}}_t(1)$. The image $xt$ can only be chosen from $E \cup \{n\}$. There are $(r+1)^r$ different mappings of $E$. Altogether we have the desired formula. \qed
\end{proof}
From Proposition~\ref{prop:sf} and Lemma~\ref{lem:Gn} we have
\begin{proposition}\label{prop:Gncard}
For $n \ge 3$, if $L$ is a suffix-free regular language with quotient complexity $n$, then its syntactic complexity $\sigma(L)$ satisfies that $\sigma(L) \le \bsf(n)$, where $\bsf(n)$ is the cardinality of $\Bsf(n)$, and it is given by Equation~(\ref{eq:g}).
\end{proposition}
Note that $\Bsf(n)$ is not a semigroup for $n \ge 4$ because $s_1 = [2,3,n,\ldots,n,n]$, $s_2 = [n,3,3,\ldots,3,n] \in \Bsf(n)$, but $s_1s_2 = [3,3,n,\ldots,n,n] \not\in \Bsf(n)$. Hence, although $\bsf(n)$ is an upper bound on the syntactic complexity of suffix-free regular languages, that bound is not tight. Our objective is to find the largest subset of $\Bsf(n)$ that is a semigroup. Let
\begin{eqnarray*}
\Vsf(n) = \{t \in \Bsf(n) &\mid& \txt{for all} i, j \in Q \txt{where} i \neq j, \\
&& \txt{we have} it = jt = n \txt{or} it \neq jt\},
\end{eqnarray*}
where $\mathbf{W}$ stands for ``witness''.
\begin{proposition}\label{prop:Pncard}
For $n \ge 3$, $\Vsf(n)$ is a semigroup contained in $\Bsf(n)$, and its cardinality is
\begin{equation*}
\vsf(n) = |\Vsf(n)| = \sum_{k = 1}^{n-1} {C^{n-1}_k}(n-1-k)!{C^{n-2}_{n-1-k}}.
\end{equation*}
\end{proposition}
\begin{proof}
We know that any $t$ is in $\Vsf(n)$ if and only if the following hold:
\begin{enumerate}
\item $it \ne 1$ for all $i \in Q$, and $nt = n$;
\item for all $i,j \in Q$, such that $i\neq j$, either $it = jt = n$ or $it \ne jt$.
\end{enumerate}
Clearly $\Vsf(n) \subseteq \Bsf(n)$. For any transformations $t_1,t_2 \in \Vsf(n)$, consider the composition $t_1t_2$. Since $1 \not\in \mathop{\mbox{rng}}(t_2)$, we have $1 \not\in \mathop{\mbox{rng}}(t_1t_2)$.
We also have $nt_1t_2 = nt_2 = n$. Pick any $i,j \in Q$ such that $i\neq j$.
Suppose $it_1t_2 \neq n$ or $jt_1t_2 \neq n$.
If $it_1t_2 = jt_1t_2$, then $it_1 = jt_1$ and thus $i = j$, a contradiction. Hence $t_1t_2 \in \Vsf(n)$, and $\Vsf(n)$ is a semigroup contained in $\Bsf(n)$.
Let $t \in \Vsf(n)$ be any transformation.
Note that $nt = n$ is fixed.
Let $Q' = Q \setminus \{n\}$, and $Q'' = Q \setminus \{1,n\}$. Suppose $k$ elements in $Q'$ are mapped to $n$ by $t$, where $0 \le k \le n-1$; then there are ${C^{n-1}_k}$ choices of these elements. For the set $D$ of the remaining $n-1-k$ elements, which must be mapped by $t$ to pairwise distinct elements of $Q''$, there are ${C^{n-2}_{n-1-k}}(n-1-k)!$ choices for the mapping $t|_D$. When $k = 0$, there is no such $t$ since $|Dt| = n-1 > n-2 = |Q''|$.
Altogether, the cardinality of $\Vsf(n)$ is
$|\Vsf(n)| = \sum_{k = 1}^{n-1} {C^{n-1}_k}(n-1-k)!{C^{n-2}_{n-1-k}}. $ \qed
\end{proof}
We now construct a generating set $\Gsf(n)$ ($\mathbf{G}$ for ``generators'') of size $n$ for $\Vsf(n)$, which will show that there exist DFA's accepting suffix-free regular languages with quotient complexity $n$ and syntactic complexity $\vsf(n)$.
\begin{proposition}\label{prop:Pgen}
When $n \ge 3$, the semigroup $\Vsf(n)$ is generated by the following set $\Gsf(n)$ of transformations of $Q$:
$\Gsf(3)=\{a,b\}$, where $a=[3,2,3]$ and $b=[2,3,3]$; $\Gsf(4)=\{a,b,c\}$, where $a= [4, 3, 2, 4]$, $b = [2, 4, 3, 4]$, $c = [2, 3, 4, 4]$; and for $n\ge 5$, $\Gsf(n)=\{a_0,\ldots,a_{n-1}\}$, where
\begin{itemize}
\item $a_0={1 \choose n}(2,3)$,
\item $a_1={1 \choose n}(2,3,\ldots,n-1)$,
\item For $2 \le i \le n - 1$, $ja_i = j+1$ for $j = 1,\ldots,i-1$, $ia_i = n$, and $ja_i = j$ for $j = i+1,\ldots,n$.
\end{itemize}
\end{proposition}
\begin{proof}
First note that $\Gsf(n)$ is a subset of $\Vsf(n)$, and so $\langle \Gsf(n) \rangle$, the semigroup generated by $\Gsf(n)$, is a subset of $\Vsf(n)$. We now show that $\Vsf(n) \subseteq \langle \Gsf(n) \rangle$.
Pick any $t$ in $\Vsf(n)$. Note that $nt = n$ is fixed.
Let $Q' = Q \setminus \{n\}$, $E_t = \{j \in Q' \mid jt = n\}$, $D_t = Q' \setminus E_t$, and $Q'' = Q \setminus \{1,n\}$. Then $D_t t \subseteq Q''$, and $|E_t| \ge 1$, since $|Q''| < |Q'|$. We prove by induction on $|E_t|$ that $t \in \langle \Gsf(n) \rangle$.
First, note that $\langle a_0,a_1 \rangle$, the semigroup generated by $\{a_0,a_1\}$, is isomorphic to the symmetric group $\mathfrak{S}_{Q''}$ by Theorem~\ref{thm:piccard}. Consider $E_t = \{i\}$ for some $i \in Q'$. Then $ia_i = it = n$. Moreover, since $D_t a_i, D_t t \subseteq Q''$, there exists $\pi \in \langle a_0, a_1 \rangle$ such that $(ja_i)\pi = jt$ for all $j \in D_t$. Then $t = a_i\pi \in \langle \Gsf(n) \rangle$.
Assume that any transformation $t \in \Vsf(n)$ with $|E_t| < k$ can be generated by $\Gsf(n)$, where $1 < k < n-1$.
Consider $t \in \Vsf(n)$ with $|E_t| = k$.
Suppose $E_t = \{e_1,\ldots,e_{k-1},e_k\}$.
Let $s\in \Vsf(n)$ be such that $E_s = \{e_1,\ldots,e_{k-1}\}$. By assumption, $s$ can be generated by $\Gsf(n)$.
Let $i = e_ks$;
then $i \in Q''$, and $e_j(s a_i) = n$ for all $1 \le j \le k$.
Moreover, we have $D_t (s a_i) \subseteq Q''$.
Thus, there exists $\pi \in \langle a_0, a_1 \rangle$ such that,
for all $d \in D_t$, $d(s a_i \pi) = d t$.
Altogether, for all $e_j \in E_t$, we have $e_j (s a_i \pi) = e_j t = n$, for all $d \in D_t$, $d (s a_i \pi) = d t$, and $n (s a_i \pi) = n t = n$. Thus $t = s a_i \pi$, and $t \in \langle \Gsf(n) \rangle$.
Therefore $\Vsf(n) = \langle \Gsf(n) \rangle$. \qed
\end{proof}
\begin{theorem}
\label{thm:DFAsf}
For $n \ge 5$, let ${\mathcal A}_n = (Q,\Sigma,\delta,1,F)$ be the DFA with alphabet $\Sigma = \{a_0,a_1,\ldots,a_{n-1}\}$, where each $a_i$ defines a transformation as in Proposition~\ref{prop:Pgen}, and $F = \{2\}$.
Then $L = L({\mathcal A}_n)$ has quotient complexity $\kappa(L) = n$, and syntactic complexity $\sigma(L) = \vsf(n)$. Moreover, $L$ is suffix-free.
\end{theorem}
\begin{proof}
First we show that all the states of ${\mathcal A}_n$ are reachable: $1$ is the initial state, state $n$ is reached by $a_1$, and for $2 \leq i \leq n-1$, state $i$ is reached by $a_i^{i-1}$.
Also, the initial state $1$ accepts $a_2$ while state $i$ rejects $a_2$ for all $i \neq 1$.
For $2 \leq i \leq n-1$, state $i$ accepts $ a_1^{n-i}$, while state $j$ rejects it, for all $j \neq i$. Also $n$ is the empty state. Thus all the states of ${\mathcal A}_n$ are distinct, and $\kappa(L) = n$.
By Proposition~\ref{prop:Pgen}, the syntactic semigroup of $L$ is $\Vsf(n)$. The syntactic complexity of $L$ is $\sigma(L) = |\Vsf(n)| = \vsf(n)$. Also, by Proposition~\ref{prop:sf}, $L$ is suffix-free. \qed
\end{proof}
As shown in Table~\ref{tab:Summary1} on p.~\pageref{table1}, the size of $\Sigma$ cannot be decreased for $n\le 5$.
\begin{theorem}\label{thm:sfsmall}
For $2 \le n \le 5$, if a suffix-free regular language $L$ has quotient complexity $\kappa(L) = n$, then its syntactic complexity satisfies that $\sigma(L) \le \vsf(n)$, and this is a tight upper bound.
\end{theorem}
\begin{proof}
By Proposition~\ref{prop:sf}, the syntactic semigroup of a suffix-free regular language $L$ is contained in $\Bsf(n)$.
For $n\in\{2,3\}$, $\vsf(n)=\bsf(n)$.
So $\vsf(n)$ is an upper bound, and it is met by the language $L = \varepsilon$ for $n = 2$ and by $L = ab^*$ for $n = 3$.
For $n=4$, we have $|\Bsf(4)| = 15$ and $|\Vsf(4)| = 13$.
Two transformations, $s_1 = [4, 2, 2, 4]$ and $s_2 = [4, 3, 3, 4]$, in $\Bsf(4)$ are such that $s_1$ conflicts with $t_1 = [3, 2, 4, 4] \in \Vsf(4)$ ($t_1s_1 = [2,2,4,4] \not\in \Bsf(4)$), and $s_2$ conflicts with $t_2 = [2,3,4,4]$ ($t_2s_2 = [3,3,4,4] \not\in \Bsf(4)$).
Thus $\sigma(L) \le 13$.
Let $L = (b \cup c)((a \cup c)b^*a)^*$; then $\kappa(L) = 4$ and $\sigma(L) = 13$. So the bound is tight.
For $n=5$, we have $|\Bsf(5)| = 115$ and $|\Vsf(5)| = 73$. Let $\Bsf(5) \setminus \Vsf(5) = \{s_1,\ldots,s_{42}\}$. For each $s_i$, we enumerated transformations in $\Vsf(5)$ using \emph{GAP} and found a unique $t_i \in \Vsf(5)$ such that the semigroup $\langle t_i, s_i \rangle$ is not contained in $\Bsf(5)$. Thus at most one transformation in each pair $\{t_i,s_i\}$ can appear in the syntactic semigroup of $L$. So we reduce the upper bound to~$73$. By Theorem~\ref{thm:DFAsf}, this bound is tight.
\end{proof}
\smallskip
For $n \ge 6$, the semigroup $\Vsf(n)$ is no longer the largest semigroup contained in $\Bsf(n)$. In the following, we define and study another semigroup $\Wsf(n)$, which is a larger semigroup contained in $\Bsf(n)$. Let
$$\Wsf(n) = \{ t \in \Bsf(n) \mid 1t = n \txt{or} it = n~~\forall~i, 2 \le i \le n-1 \}.$$
Note that, we are interested only in situations where $n \ge 6$, although some statements also hold for smaller $n$.
\begin{proposition}\label{prop:Wsf}
For $n \ge 6$, the set $\Wsf(n)$ is a semigroup contained in $\Bsf(n)$, and its cardinality is
$${\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n) = |\Wsf(n)| = (n-1)^{n-2} + (n-2).$$
\end{proposition}
\begin{proof}
Pick any $t_1,t_2$ in $\Wsf(n)$. If $1t_1 = n$, then $1(t_1t_2) = n$ and $t_1t_2 \in \Wsf(n)$. If $1t_1 \neq n$, then, for all $i \in \{2,\ldots,n-1\}$, $it_1 = n$ and $i(t_1t_2) = n$; so $t_1t_2 \in \Wsf(n)$ as well. Hence $\Wsf(n)$ is a semigroup contained in $\Bsf(n)$.
For any $t \in \Wsf(n)$, $nt = n$ is fixed. There are two possible cases:
\begin{enumerate}
\item $1t = n$: For each $i \in \{2,\ldots,n-1\}$, $it$ can be chosen from $\{2,\ldots,n\}$. Then there are $(n-1)^{n-2}$ different $t$'s in this case.
\item $1t \neq n$: Now $1t$ can be chosen from $\{2,\ldots,n-1\}$. For each $i \in \{2,\ldots,n-1\}$, $it = n$ is fixed. There are $n-2$ different $t$'s in this case.
\end{enumerate}
Therefore ${\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n) = (n-1)^{n-2} + (n-2)$. \qed
\end{proof}
\begin{proposition}\label{prop:Wsfgen}
For $n \ge 6$, the semigroup $\Wsf(n)$ is generated by the set $\mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) = \{a_1,a_2,a_3,b_1,\ldots,b_{n-2},c\}$ of transformations, where
\begin{enumerate}
\item $a_1 = {1 \choose n}(2,\ldots,n-1)$, $a_2 = {1 \choose n}(2,3)$, $a_3 = {1 \choose n}{n-1 \choose 2}$;
\item For $1 \le i \le n-2$, $b_i = {1 \choose n}{i+1 \choose n}$;
\item $c = {Q \setminus \{1\} \choose n}{1 \choose 2} = [2,n,\ldots,n]$.
\end{enumerate}
\end{proposition}
\begin{proof}
Clearly $\mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) \subseteq \Wsf(n)$, and $\langle \mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) \rangle \subseteq \Wsf(n)$. We show in the following that $\Wsf(n) \subseteq \langle \mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) \rangle$.
Let $Q' = \{2,\ldots,n-1\}$. By Theorem~\ref{thm:salomaa}, $a_1,a_2$ and $a_3$ together generate the semigroup $$\mathbf{Y} = \{t \in \Wsf(n) \mid \txt{for all} i \in Q', it \in Q'\},$$ which is isomorphic to ${\mathcal T}_{Q'}$ and is contained in $\Wsf(n)$. Next, consider any $t \in \Wsf(n) \setminus \mathbf{Y}$. We have two cases:
\begin{enumerate}
\item $1t = n$: Let $E_t = \{ i \in Q' \mid it = n \}$. Since $t \not\in \mathbf{Y}$, $E_t \neq \emptyset$. Suppose $E_t = \{i_1,\ldots,i_k\}$, for some $1 \le k \le n-2$. Then there exists $t' \in \mathbf{Y}$ such that, for all $i \not\in E_t$, $it' = it$. Let $s = b_{i_1-1} \cdots b_{i_k-1}$. Note that $E_ts = \{n\}$, and, for all $i \not\in E_t$, $i(t's) = (it')s = it$. So $t = t's \in \langle \mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) \rangle$.
\item $1t \neq n$: If $1t = 2$, then $t = c$. Otherwise, $1t \in \{3,\ldots,n-1\} \subseteq Q'$, and we know from the above case that there exists $t' \in \mathbf{G}^{\ge 6}_{\mathrm{sf}}(n)$ such that $2t' = 1t$. Then $1(ct') = 1t$, and $i(ct') = (ic)t' = n = it$, for all $i \in Q'$. Hence $t = ct'~\in~\langle \mathbf{G}^{\ge 6}_{\mathrm{sf}}(n) \rangle$.
\end{enumerate}
Therefore $\langle a_1,a_2,a_3,b_1,\ldots,b_{n-2},c \rangle = \Wsf(n)$. \qed
\end{proof}
\begin{theorem}\label{thm:wsfaut}
For $n \ge 6$, let ${\mathcal A}'_n = (Q, \Sigma, \delta, 1, F)$ be the DFA with alphabet $\Sigma = \{a_1,a_2,a_3,b_1,\ldots,b_{n-2},c\}$ of size $n+2$, where each letter defines a transformation as in Proposition~\ref{prop:Wsfgen}, and $F = \{2\}$. Then $L' = L({\mathcal A}'_n)$ has quotient complexity $\kappa(L') = n$ and syntactic complexity $\sigma(L') = {\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n)$.
\end{theorem}
\begin{proof}
First we show that $\kappa(L') = n$. From the initial state, we can reach state $2$ by $c$ and state $n$ by $a_1$. From state $2$ we can reach state $i$, $3 \le i \le n-1$, by $a_1^{i-1}$. So all the states in $Q$ are reachable. Now, the initial state accepts $c$, but all other states reject it. For $2 \le i \le n-2$, state $i$ accepts $a_1^{n-i}$, while all other states reject it. State $n$ is the empty state, which rejects all words. Thus all the states in $Q$ are distinct.
By Proposition~\ref{prop:Wsfgen}, the syntactic semigroup of $L'$ is $\Wsf(n)$, and $\sigma(L') = {\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n)$. Also $L'$ is suffix-free by Proposition~\ref{prop:sf}. \qed
\end{proof}
We know that the upper bound on the syntactic complexity of suffix-free regular languages is achieved by the largest semigroup contained in $\Bsf(n)$. We conjecture that $\Wsf(n)$ is such a semigroup.
\medskip
\begin{conjecture}[Suffix-Free Regular Languages]
\label{con:sf}
If $L$ is a suffix-free regular language with $\kappa(L) = n \ge 6$, then $\sigma(L) \le {\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n)$ and this is a tight bound.
\end{conjecture}
We prove the conjecture for $n = 6$:
\begin{proof}
For $n = 6$, $|\Bsf(6)| = 1169$ and $|\Wsf(6)| = 629$. Let $\{s_1,\ldots,s_{540}\} = \Bsf(6) \setminus \Wsf(6)$. For each $i$, we enumerated transformations in $\Wsf(6)$ using \emph{GAP} and found a unique $t_i \in \Wsf(6)$ such that $\langle t_i, s_i \rangle$ is not contained in $\Bsf(6)$. As in the proof of Theorem~\ref{thm:sfsmall}, for each $i$, at most one transformation in $\{t_i,s_i\}$ can appear in the syntactic semigroup of $L$. Then we can reduce the upper bound to $629$. This bound is met by the language $L'$ in Theorem~\ref{thm:wsfaut}; so it is tight. \qed
\end{proof}
\section{Bifix-Free Regular Languages}\label{sec:bf}
Let $L$ be a regular bifix-free language with $\kappa(L) = n$. From Sections~\ref{sec:pf} and~\ref{sec:sf} we have:
\begin{enumerate}
\item $L$ has $\varepsilon$ as a quotient, and this is the only accepting quotient;
\item $L$ has $\emptyset$ as a quotient;
\item $L$ as a quotient is uniquely reachable.
\end{enumerate}
Let ${\mathcal A}$ be the quotient DFA of $L$, with $Q$ as the set of states. We assume that $1$ is the initial state, $n-1$ corresponds to the quotient $\varepsilon$, and $n$ is the empty state. Consider the set
$$\Bbf(n) = \{ t \in \Bsf(n) \mid (n-1) t = n \}.$$
The following is an observation similar to Proposition~\ref{prop:sf}.
\begin{proposition}\label{prop:bf}
If $L$ is a regular language with quotient complexity $n$ and syntactic semigroup $T_L$, then the following hold:
\begin{enumerate}
\item If $L$ is bifix-free, then $T_L$ is a subset of $\Bbf(n)$.
\item If $\varepsilon$ is the only accepting quotient of $L$, and $T_L \subseteq \Bbf(n)$, then $L$ is bifix-free.
\end{enumerate}
\end{proposition}
\begin{proof}\mbox{}
1. Since $L$ is suffix-free, $T_L \subseteq \Bsf(n)$. Since $L$ is also prefix-free, it has $\varepsilon$ and $\emptyset$ as quotients. By assumption, $n-1 \in Q$ corresponds to the quotient $\varepsilon$. Thus for any $t \in T_L$, $(n-1) t = n$, and so $T_L \subseteq \Bbf(n)$.
2. Since $\varepsilon$ is the only accepting quotient of $L$, $L$ is prefix-free, and $L$ has the empty quotient. Since $T_L \subseteq \Bbf(n) \subseteq \Bsf(n)$, $L$ is suffix-free by Proposition~\ref{prop:sf}. Therefore $L$ is bifix-free. \qed
\end{proof}
\begin{lemma}\label{lem:Hn}
For $n \ge 3$, we have $|\Bbf(n)| = M_n + N_n$, where
\begin{eqnarray}
\label{eq:H1}
M_n &=& \sum_{k=1}^{n-2} C^{n-3}_{k-1} (k-1)! \sum_{\begin{subarray}{c}
r_2 + \cdots + r_k + r \\
= n-k-2
\end{subarray}}
C^{n-k-2}_{r_2,\ldots,r_k,r} (r+1)^r \prod_{j=2}^k S'_{r_j+1}(j-1), \\
\label{eq:H2}
N_n &=& \sum_{k=0}^{n-3} C^{n-3}_k k! \sum_{\begin{subarray}{c}
r_2 + \cdots + r_k + r \\
= n-k-3
\end{subarray}}
C^{n-k-3}_{r_2,\ldots,r_k,r} (r+2)^r \prod_{j=2}^k S'_{r_j+1}(j-1) .
\end{eqnarray}
\end{lemma}
\begin{proof}
Let $t$ be any transformation in $\Bbf(n)$. Suppose ${\mathfrak{s}}_t(1) = 1,1t,\ldots,1t^k,n$, where $0 \le k \le n-2$. For $2 \le j \le k$, suppose tree $T_t(j)$ contains $r_j+1$ nodes, for some $r_j \ge 0$; then there are $S'_{r_j+1}(j-1)$ different trees $T_t(j)$. Let $E$ be the set of elements of $Q$ that are not in any tree $T_t(j)$ nor in the principal sequence~${\mathfrak{s}}_t(1)$. Then there are two cases:
\begin{enumerate}
\item $n-1 \in {\mathfrak{s}}_t(1)$: Since $(n-1)t = n$, we must have $1t^k = n-1$, and $k \ge 1$. So there are $C^{n-3}_{k-1} (k-1)!$ different ${\mathfrak{s}}_t(1)$. Let $r = |E| = (n-k-2) - (r_2 + \cdots + r_k)$. Then there are $C^{n-k-2}_{r_2,\ldots,r_k,r}$ tuples $(r_2,\ldots,r_k,r)$. For any $x \in E$, its image $xt$ can be chosen from $E \cup \{n\}$. Then the number of transformations $t$ in this case is $M_n$.
\item $n-1 \not\in {\mathfrak{s}}_t(1)$: Then $k \le n-3$, and there are $C^{n-3}_k k!$ different ${\mathfrak{s}}_t(1)$. Note that $n-1 \in E$, and $(n-1)t = n$ is fixed. Let $r = |E \setminus \{n-1\}| = (n-k-3) - (r_2 + \cdots + r_k)$. Then there are $C^{n-k-3}_{r_2,\ldots,r_k,r}$ tuples $(r_2,\ldots,r_k,r)$. For any $x \in E \setminus \{n-1\}$, $xt$ can be chosen from $E \cup \{n\}$. Thus the number of transformations $t$ in this case is $N_n$.
\end{enumerate}
Altogether we have the desired formula. \qed
\end{proof}
Let $\bbf(n) = |\Bbf(n)|$. From Proposition~\ref{prop:bf} and Lemma~\ref{lem:Hn} we have
\begin{proposition}\label{prop:Hncard}
For $n \ge 3$, if $L$ is a bifix-free regular language with quotient complexity $n$, then its syntactic complexity $\sigma(L)$ satisfies that $\sigma(L) \le \bbf(n)$, where $\bbf(n)$ is the cardinality of $\Bbf(n)$ as in Lemma~\ref{lem:Hn}.
\end{proposition}
For $2 \le n \le 4$, the set $\Bbf(n)$ is a semigroup. But for $n \ge 5$, it is not a semigroup because $s_1 = [2,3,n,\ldots,n,n]$, $s_2 = [n,3,3,n,\ldots,n,n] \in \Bbf(n)$ while $s_1s_2 = [3,3,n,\ldots,n,n] \not\in \Bbf(n)$. Hence $\bbf(n)$ is not a tight upper bound on the syntactic complexity of bifix-free regular languages in general. We look for a large semigroup contained in $\Bbf(n)$ that can be the syntactic semigroup of a bifix-free regular language. Let
\begin{eqnarray*}
\Vbf(n) = \{ t \in \Bbf(n) &\mid& \txt{for all} i, j \in Q \txt{where} i \neq j, \\
&& \txt{we have} it = jt = n \txt{or} it \neq jt \}.
\end{eqnarray*}
(The reason for using the superscript $\le 5$ will be made clear in Theorem~\ref{thm:bfsmall}.)
\begin{proposition}\label{prop:Vncard}
For $n \ge 3$, $\Vbf(n)$ is a semigroup contained in $\Bbf(n)$ with cardinality
\begin{equation*}
\vbf(n) = |\Vbf(n)| = \sum_{k=0}^{n-2} \big( C^{n-2}_k \big)^2 (n-2-k)!
\end{equation*}
\end{proposition}
\begin{proof}
First, note that $\Vbf(n) = \Vsf(n) \cap \Bbf(n)$, and that $\Vsf(n)$ is a semigroup contained in $\Bsf(n)$ by Proposition~\ref{prop:Pncard}. For any $t_1,t_2 \in \Vbf(n)$, we have $t_1t_2 \in \Vsf(n)$, and $(n-1)t_1t_2 = nt_2 = n$; so $t_1t_2 \in \Bbf(n)$. Then $t_1t_2 \in \Vbf(n)$, and $\Vbf(n)$ is a semigroup contained in~$\Bbf(n)$.
Pick any $t \in \Vbf(n)$. Note that $(n-1)t = n$ and $nt = n$ are fixed, and $1 \not\in \mathop{\mbox{rng}}(t)$. Let $Q' = Q \setminus \{n-1,n\}$, $E = \{i \in Q' \mid it = n\}$, and $D = Q' \setminus E$. Suppose $|E| = k$, where $0 \leq k \leq n-2$; then there are ${C^{n-2}_k}$ choices of $E$. Elements of $D$ are mapped to pairwise different elements of $Q \setminus \{1,n\}$; then there are ${C^{n-2}_{n-2-k}}(n-2-k)!$ different mappings $t|_D$. Altogether, we have
$ |\Vbf(n)| = \sum_{k=0}^{n-2} \big( C^{n-2}_k \big)^2 (n-2-k)! $ \qed
\end{proof}
\begin{proposition}\label{prop:bfgen}
For $n \ge 3$, let $Q' = Q \setminus \{n-1, n\}$ and $Q'' = Q \setminus \{1,n\}$. Then the semigroup $\Vbf(n)$ is
generated by
$$\Gbf(n) = \{t \in \Vbf(n) \mid Q't = Q'' \txt{and} it \neq jt \txt{for all} i,j \in Q'\}.$$
\end{proposition}
\begin{proof}\label{proof:bfgen}
We want to show that $\Vbf(n) = \langle \Gbf(n) \rangle$. Since $\Gbf(n) \subseteq \Vbf(n)$, we have $\langle \Gbf(n) \rangle \subseteq \Vbf(n)$.
Let $t \in \Vbf(n)$. By definition, $(n-1)t = nt = n$. Let $E_{t} = \{i \in Q' \mid it = n\}$. If $E_{t} = \emptyset$, then $t \in \Gbf(n)$;
otherwise, there exists $x \in Q''$ such that $x \not\in \mathop{\mbox{rng}}(t)$. We prove by induction on $|E_{t}|$ that $t \in \langle \Gbf(n) \rangle$.
First note that, for all $t \in \Gbf(n)$, $t|_{Q'}$ is an injective mapping from $Q'$ to $Q''$. Consider $E_{t} = \{i\}$ for some $i \in Q'$.
Since $|E_{t}| = 1$, $\mathop{\mbox{rng}}(t) \cup \{x\} = Q''$. Let $t_1,t_2 \in \Gbf(n)$ be defined by
\begin{enumerate}
\item $jt_1 = j+1$ for $j = 1,\ldots,i-1$, $it_1 = n-1$, $jt_1 = j$ for $j = i+1,\ldots,n-2$,
\item $1t_2 = x$, $jt_2 = (j-1)t$ for $j = 2,\ldots,i$, $jt_2 = jt$ for $j = i+1,\ldots,n-2$.
\end{enumerate}
Then $t_1t_2 = t$, and $t \in \langle \Gbf(n) \rangle$.
\goodbreak
Assume that any transformation $t \in \Vbf(n)$ with $|E_{t}| < k$ can be generated by $\Gbf(n)$, where $1 < k < n-2$.
Consider $t \in \Vbf(n)$ with $|E_{t}| = k$.
Suppose $E_{t} = \{e_1,\ldots,e_{k-1},e_k\}$, and let $D_{t} = Q' \setminus E_{t} = \{d_1,\ldots,d_l\}$, where $l = n - 2 - k$.
By assumption, all $s \in \Vbf(n)$ with $|E_{s}| = k-1$ can be generated by $\Gbf(n)$.
Let $s$ be such that $E_{s} = \{1,\ldots,k-1\}$; then $1s = \cdots = (k-1)s = n$. In addition, let $ks = x$, and let $(k+j)s = d_jt$ for $j = 1,\ldots,l$. Let $t' \in \Gbf(n)$ be such that $e_jt' = j$ for $j = 1,\ldots,k-1$, $kt' = n-1$, and $d_jt' = k+j$ for $j = 1,\ldots,l$. Then $t's = t$, and $t \in \langle \Gbf(n) \rangle$.
Therefore, $\Vbf(n) = \langle \Gbf(n) \rangle$. \qed
\end{proof}
\begin{theorem}\label{thm:bfaut}
For $n \ge 3$, let ${\mathcal A}_n = (Q,\Sigma,\delta,1,F)$ be the DFA
with alphabet $\Sigma$ of size $(n-2)!$, where each $a \in \Sigma$ defines a distinct transformation $t_a \in \Gbf(n)$, and $F = \{n-1\}$.
Then $L = L({\mathcal A}_n)$ has quotient complexity $\kappa(L) = n$, and syntactic complexity $\sigma(L) = \vbf(n)$.
Moreover, $L$ is bifix-free.
\end{theorem}
\begin{proof}\label{proof:bfaut}
We first show that all the states of ${\mathcal A}_n$ are reachable.
Note that there exists $a \in \Sigma$ such that $t_a = [2,3,\ldots,n-1,n,n] \in \Gbf(n)$.
State $1 \in Q$ is the initial state, and
$a^{i-1}$ reaches state $i \in Q$ for $i = 2,\ldots,n$.
Furthermore, for $1 \leq i \leq n-1$, state $i$ accepts $ a^{n-1-j}$, while for $j \neq i$, state $j$ rejects it. Also, $n$ is the empty state. Thus all the states of ${\mathcal A}_n$ are distinct, and $\kappa(L) = n$.
By Proposition~\ref{prop:bfgen}, the syntactic semigroup of $L$ is $\Vbf(n)$. Hence the syntactic complexity of $L$ is $\sigma(L) = \vbf(n)$. By Proposition~\ref{prop:bf}, $L$ is bifix-free. \qed
\end{proof}
\begin{theorem}\label{thm:bfsmall}
For $2 \le n \le 5$, if a bifix-free regular language $L$ has quotient complexity $\kappa(L) = n$, then $\sigma(L) \le \vbf(n)$, and this bound is tight.
\end{theorem}
\begin{proof}
We know by Proposition~\ref{prop:bf} that the upper bound on the syntactic complexity of bifix-free regular languages is reached by the largest semigroup contained in $\Bbf(n)$. Since $\vbf(n) = \bbf(n)$ for $n = 2$, $3$, and $4$, $\vbf(n)$ is an upper bound, and it is tight by Theorem~\ref{thm:bfaut}.
For $n = 5$, we have $\bbf(5) = |\Bbf(5)| = 41$, and $\vbf(5) = |\Vbf(5)| = 34$. Let $\Bbf(5) \setminus \Vbf(5) = \{\tau_1,\ldots,\tau_7\}$. We found for each $\tau_i$ a unique $t_i \in \Vbf(5)$ such that the semigroup $\langle \tau_i, t_i \rangle$ is not a subset of $\Bbf(5)$:
$$\begin{array}{ll}
\tau_1 = [ 2, 4, 4, 5, 5 ], \quad & \quad t_1 = [ 3, 4, 2, 5, 5 ]; \\
\tau_2 = [ 3, 4, 4, 5, 5 ], \quad & \quad t_2 = [ 3, 5, 2, 5, 5 ]; \\
\tau_3 = [ 4, 2, 2, 5, 5 ], \quad & \quad t_3 = [ 2, 4, 3, 5, 5 ]; \\
\tau_4 = [ 4, 3, 3, 5, 5 ], \quad & \quad t_4 = [ 2, 5, 3, 5, 5 ]; \\
\tau_5 = [ 5, 2, 2, 5, 5 ], \quad & \quad t_5 = [ 3, 2, 4, 5, 5 ]; \\
\tau_6 = [ 5, 3, 3, 5, 5 ], \quad & \quad t_6 = [ 2, 3, 4, 5, 5 ]; \\
\tau_7 = [ 5, 4, 4, 5, 5 ], \quad & \quad t_7 = [ 3, 2, 5, 5, 5 ].
\end{array}$$
Since $\langle \tau_i, t_i \rangle \subseteq T_L$, if both $\tau_i$ and $t_i$ are in $T_L$,
then $T_L \not\subseteq \Bbf(5)$, and $L$ is not bifix-free by Proposition~\ref{prop:bf}. Thus, for $1 \le i \le 7$, at most one of $\tau_i$ and $t_i$ can appear in $T_L$, and $|T_L| \le 34$. Since $|\Vbf(5)| = 34$ and $\Vbf(5)$ is a semigroup, we have $\sigma(L) \le 34 = \vbf(5)$ as the upper bound for $n = 5$. This bound is reached by the DFA ${\mathcal A}_5$ in Theorem~\ref{thm:bfaut}. \qed
\end{proof}
For $n \ge 6$, the semigroup $\Vbf(n)$ is no longer the largest semigroup contained in $\Bbf(n)$. We find another large semigroup $\Wbf(n)$ suitable for bifix-free regular languages. Let
\begin{eqnarray*}
\mathbf{U}^1_n &=& \{t \in \Bbf(n) \mid 1t = n \}, \\
\mathbf{U}^2_n &=& \{t \in \Bbf(n) \mid 1t = n-1\}, \\
\mathbf{U}^3_n &=& \{t \in \Bbf(n) \mid 1t \not\in \{n, n-1\}, \txt{and} it \in \{n-1, n\} \txt{for all} i \neq 1\},
\end{eqnarray*}
and let $\Wbf(n) = \mathbf{U}^1_n \cup \mathbf{U}^2_n \cup \mathbf{U}^3_n$.
When $2 \le n \le 4$, we have $\Wbf(n)~=~\Vbf(n)$, and these cases were already discussed. So we are only interested in larger values~of~$n$.
\begin{proposition}\label{prop:Rcard}
For $n \ge 5$, $\Wbf(n)$ is a semigroup contained in $\Bbf(n)$ with cardinality
$$ \wbf(n) = |\Wbf(n)| = (n-1)^{n-3} + (n-2)^{n-3} + (n-3)2^{n-3}. $$
\end{proposition}
\begin{proof}
First we show that $\mathbf{U}^1_n$ is a semigroup. For any $t_1,t_1' \in \mathbf{U}^1_n$, since $1(t_1t_1') = (1t_1)t_1' = nt_1' = n$, we have $t_1t_1' \in \mathbf{U}^1_n$. Next, let $t_2 \in \mathbf{U}^2_n$ and $t \in \mathbf{U}^1_n \cup \mathbf{U}^2_n$. If $t \in \mathbf{U}^1_n$, then $1(t_2t) = (n-1)t = n$ and $1(tt_2) = nt_2 = n$; so $t_2t, tt_2 \in \mathbf{U}^1_n$. If $t \in \mathbf{U}^2_n$, then $1(t_2t) = (n-1)t = n$ and $1(tt_2) = (n-1)t_2 = n$; so $t_2t, tt_2 \in \mathbf{U}^1_n$ as well. Thus $\mathbf{U}^1_n \cup \mathbf{U}^2_n$ is also a semigroup. For any $t_3 \in \mathbf{U}^3_n$ and $t' \in \Wbf(n)$, since $it_3 \in \{n-1,n\}$ for all $i \neq 1$, and $(n-1)t' = nt' = n$, we have $i(t_3t') = n$, and $t_3t' \in \Wbf(n)$. Also $1(t't_3) = (1t')t_3 \in \{n-1,n\}$, so $t't_3 \in \mathbf{U}^1_n \cup \mathbf{U}^2_n$. Hence $\Wbf(n)$ is a semigroup contained in $\Bbf(n)$.
Note that $\mathbf{U}^1_n$, $\mathbf{U}^2_n$, and $\mathbf{U}^3_n$ are pairwise disjoint. For any $t \in \Wbf(n)$, there are three cases:
\begin{enumerate}
\item $t \in \mathbf{U}^1_n$: For any $i \not\in \{1,n-1,n\}$, $it$ can be chosen from $Q \setminus \{1\}$. Then $|\mathbf{U}^1_n| = (n-1)^{n-3}$;
\item $t \in \mathbf{U}^2_n$: For any $i \not\in \{1,n-1,n\}$, $it$ can be chosen from $Q \setminus \{1,n-1\}$. Then $|\mathbf{U}^2_n| = (n-2)^{n-3}$;
\item $t \in \mathbf{U}^3_n$: Now, $1t$ can be chosen from $Q \setminus \{1,n-1,n\}$. For any $i \not\in \{1,n-1,n\}$, $it$ has two choices: $it = n-1$ or $n$. Then $|\mathbf{U}^3_n| = (n-3)2^{n-3}$.
\end{enumerate}
Therefore we have $|\Wbf(n)| = (n-1)^{n-3} + (n-2)^{n-3} + (n-3)2^{n-3}$. \qed
\end{proof}
The next proposition describes a generating set of $\Wbf(n)$.
\begin{proposition}\label{prop:Rgen}
For $n \ge 5$, the semigroup $\Wbf(n)$ is generated by $\mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) = \{a_1, a_2, a_3, b_1, \ldots, b_{n-3}, c_1, \ldots, c_m, d_1, \ldots, d_l\}$, where $m = (n-2)^{n-3}-1$ and $l = (n-3)(2^{n-3} - 1)$, and
\begin{enumerate}
\item $a_1 = {1 \choose n}{n-1 \choose n}(2,\ldots,n-2)$, $a_2 = {1 \choose n}{n-1 \choose n}(2,3)$, $a_3 = {1 \choose n}{n-1 \choose n}{n-2 \choose 2}$;
\item For $1 \le i \le n-3$, $b_i = {1 \choose n}{n-1 \choose n}{i+1 \choose n-1}$;
\item Each $c_i$ defines a distinct transformation in $\mathbf{U}^2_n$ other than $[n-1,n,\ldots,n,n]$;
\item Each $d_i$ defines a distinct transformation in $\mathbf{U}^3_n$ other than $[j,n,\ldots,n,n]$ for all $j \in \{2,\ldots,n-2\}$.
\end{enumerate}
\end{proposition}
\begin{proof}
Since $\mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \subseteq \Wbf(n)$, we have $\langle \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \rangle \subseteq \Wbf(n)$. It remains to be shown that $\Wbf(n) \subseteq \langle \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \rangle$. Let $Q' = Q \setminus \{1,n-1,n\}$.
\begin{enumerate}
\item First consider $\mathbf{U}^1_n$. By Theorem~\ref{thm:salomaa}, $a_1,a_2$ and $a_3$ together generate the semigroup $$\mathbf{Y}' = \{t \in \mathbf{U}^1_n \mid \txt{for all} i \in Q', it \in Q'\},$$ which is contained in $\mathbf{U}^1_n$. For any $t \in \mathbf{U}^1_n \setminus \mathbf{Y}'$, let $E_t = \{ i \in Q \mid it = n-1 \}$; then $E_t \neq \emptyset$. Suppose $E_t = \{i_1,\ldots,i_k\}$, where $1 \le k \le n-3$. Then there exists $t' \in \mathbf{Y}'$ such that, for all $i \not\in E_t$, $it' = it$. Let $s = b_{i_1-1} \cdots b_{i_k-1}$. Note that $E_ts = \{n-1\}$, and, for all $i \not\in E_t$, $i(t's) = (it')s = it$. So $t's = t$, and $\langle a_1,a_2,a_3,b_1,\ldots,b_{n-3} \rangle = \mathbf{U}^1_n$.
\item Next, the transformations that are in $\mathbf{U}^2_n \cup \mathbf{U}^3_n$ but not in $\mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$ are $t_i = [i,n,\ldots,n,n]$, where $2 \le i \le n-1$. Note that $d = {1 \choose 2}{n-1 \choose n}{Q' \choose n-1} \in \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$, and, for each $i \in \{2,\ldots,n-1\}$, $s_i = {1 \choose n}{n-1 \choose n}{2 \choose i} \in \mathbf{U}^1_n$. Then $t_i = ds_i \in \langle \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \rangle$, and $\mathbf{U}^2_n \cup \mathbf{U}^3_n \subseteq \langle \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \rangle$.
\end{enumerate}
Therefore $\Wbf(n) = \langle \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n) \rangle$. \qed
\end{proof}
\begin{theorem}\label{thm:bfaut1}
For $n \ge 5$, let ${\mathcal A}'_n = (Q, \Sigma, \delta, 1, F)$ be the DFA with alphabet $\Sigma$ of size $(n-2)^{n-3} + (n-3)2^{n-3}+2$, where each letter defines a transformation as in Proposition~\ref{prop:Rgen}, and $F = \{n-1\}$. Then $L' = L({\mathcal A}'_n)$ has quotient complexity $\kappa(L') = n$, and syntactic complexity $\sigma(L') = \wbf(n)$. Moreover, $L'$ is bifix-free.
\end{theorem}
\begin{proof}
First, for all $i \in Q \setminus \{1\}$, there exists $a \in \Sigma$ such that $t_a = [i, n, \ldots, n, n] \in \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$, and state $i$ is reachable by $a$. So all the states in $Q$ are reachable. Next, there exist $b, c \in \Sigma$ such that $t_b = [n-1, n, \ldots, n, n] \in \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$ and $t_c = [n, 3, 4, \ldots, n, n] \in \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$. The initial state accepts $b$, while all other states reject it. For $2 \le i \le n-2$, state $i$ accepts $b^{n-i-1}$, while all other states reject it. Also, state $n-1$ is the only accepting state, and state $n$ is the empty state. Then all the states in $Q$ are distinct, and $\kappa(L') = n$.
By Proposition~\ref{prop:Rgen}, the syntactic semigroup of $L'$ is $\Wbf(n)$; so $\sigma(L')=\wbf(n)$. By Proposition~\ref{prop:bf}, $L'$ is bifix-free. \qed
\end{proof}
\smallskip
\begin{conjecture}[Bifix-Free Regular Languages]
\label{con:bf}
If $L$ is a bifix-free regular language with $\kappa(L) = n \ge 6$, then $\sigma(L) \le \wbf(n)$ and this is a tight bound.
\end{conjecture}
\smallskip
The conjecture holds for $n=6$ as we now show:
\begin{proof}
When $n = 6$, $|\Bbf(6)| = 339$ and $|\Vbf(6)| = 213$. There are 126 transformations $\tau_1,\ldots,\tau_{126}$ in $\Bbf(6) \setminus \Vbf(6)$. For each $\tau_i$, we enumerated transformations in $\Wbf(6)$ using \emph{GAP} and found a unique $t_i \in \Vbf(6)$ such that $\langle t_i, \tau_i \rangle \not\subseteq \Bbf(6)$. Thus, for each $i$, at most one of $t_i$ and $\tau_i$ can appear in the syntactic semigroup $T_L$ of $L$. So we further lower the bound to $\sigma(L) \le 213$. This bound is reached by the DFA ${\mathcal A}'_6$ in Theorem~\ref{thm:bfaut1}; so it is a tight upper bound for~$n~=~6$. \qed
\end{proof}
\section{Factor-Free Regular Languages}\label{sec:ff}
Let $L$ be a factor-free regular language with $\kappa(L) = n$. Since factor-free regular languages are also bifix-free, $L$ as a quotient is uniquely reachable, $\varepsilon$ is the only accepting quotient of $L$, and $L$ also has the empty quotient. As in Section~\ref{sec:bf}, we assume that $Q$ is the set of states of quotient DFA of $L$, in which $1$ is the initial state, and states $n-1$ and $n$ correspond to the quotients $\varepsilon$~and~$\emptyset$, respectively. Let
$$\mathbf{B}_{\mathrm{ff}}(n) = \{t \in \Bbf(n) \mid \txt{for all} j \ge 1, 1t^j = n - 1 \Rightarrow it^j = n ~~\forall~i, 1 < i < n-1 \}.$$
We first have the following observation:
\begin{proposition}\label{prop:ff}
If $L$ is a regular language with quotient complexity $n$ and syntactic semigroup $T_L$, then the following hold:
\begin{enumerate}
\item If $L$ is factor-free, then $T_L$ is a subset of $\mathbf{B}_{\mathrm{ff}}(n)$.
\item If $\varepsilon$ is the only accepting quotient of $L$, and $T_L \subseteq \mathbf{B}_{\mathrm{ff}}(n)$, then $L$ is factor-free.
\end{enumerate}
\end{proposition}
\begin{proof}
1. Assume $L$ is factor-free. Then $L$ is bifix-free, and $T_L \subseteq \Bbf(n)$ by Proposition~\ref{prop:bf}. For any transformation $t_w \in T_L$ performed by some non-empty word $w$, if $1t^j_w = n - 1$ for some $j \ge 1$, then $w^j \in L$. If we also have $it^j_w \neq n$ for some $i \in Q \setminus \{1\}$, then $i \not\in \{n - 1, n\}$ as $(n-1)t = nt = n$ for all $t \in \mathbf{B}_{\mathrm{ff}}(n)$. Thus there exist non-empty words $u$ and $v$ such that state $i$ is reachable by $u$, and state $i(t^j_w)$ accepts $v$. So $uw^jv \in L$, which is a contradiction. Hence $T_L \subseteq \mathbf{B}_{\mathrm{ff}}(n)$.
2. Since $\varepsilon$ is the only accepting state and $\mathbf{B}_{\mathrm{ff}}(n) \subseteq \Bbf(n)$, $L$ is bifix-free by Proposition~\ref{prop:bf}. If $L$ is not factor-free, then there exist non-empty words $u, v$ and $w$ such that $w, uwv \in L$. Thus $1t_w = n - 1$, and $1t_{uwv} = 1(t_ut_wt_v) = n - 1$. Since $L$ is bifix-free, $1t_u \neq 1$ and $nt_v = n$; thus $(1t_u)t_w \neq n$, which contradicts the assumption that $t_w \in T_L \subseteq \mathbf{B}_{\mathrm{ff}}(n)$. Therefore $L$ is bifix-free. \qed
\end{proof}
The properties of suffix- and bifix-free regular languages still apply to factor-free regular languages. Moreover, we have
\begin{lemma}\label{lem:ffseq}
For all $t \in \mathbf{B}_{\mathrm{ff}}(n)$ and $i \not\in {\mathfrak{s}}_t(1)$, if $n-1 \in {\mathfrak{s}}_t(1)$, then $n \in {\mathfrak{s}}_t(i)$.
\end{lemma}
\begin{proof}
Suppose $n-1 = 1t^k \in {\mathfrak{s}}_t(1)$ for some $k \ge 1$. If $n \not\in {\mathfrak{s}}_t(i)$, then for all $j \ge 1$, $it^j \neq n$. In particular, $it^k \neq n$, which contradicts the definition of $\mathbf{B}_{\mathrm{ff}}(n)$. Therefore $n \in {\mathfrak{s}}_t(i)$. \qed
\end{proof}
\begin{lemma}\label{lem:Fn}
For $n \ge 3$, we have $|\mathbf{B}_{\mathrm{ff}}(n)| = N_n + O_n$, where
\begin{eqnarray*}
O_n &=& 1 + \sum_{k=2}^{n-2} C^{n-3}_{k-1} (k-1)! \sum_{\begin{subarray}{c}
r_2 + \cdots + r_k + r \\
= n-k-2
\end{subarray}}
C^{n-k-2}_{r_2,\ldots,r_k,r} S'_{r+1}(k) \prod_{j=2}^k S'_{r_j+1}(j-1),
\end{eqnarray*}
and $N_n$ as given in Equation~(\ref{eq:H2}).
\end{lemma}
\begin{proof}
Let $t \in \mathbf{B}_{\mathrm{ff}}(n)$ be any transformation. Suppose ${\mathfrak{s}}_t(1) = 1,1t,\ldots,1t^k,n$, where $0 \le k \le n-2$. Then there are two cases:
\begin{enumerate}
\item $n-1 \in {\mathfrak{s}}_t(1)$. Since $(n-1)t = n$, we have $n-1 = 1t^k$, and $k \ge 1$. If $k = 1$, then $1t = n-1$, and $it = n$ for all $i \neq 1$; such a $t$ is unique. Consider $k \ge 2$. There are $C^{n-2}_{k-1} (k-1)!$ different ${\mathfrak{s}}_t(1)$. For $2 \le j \le k$, suppose there are $r_j + 1$ nodes in tree $T_t(j)$; then there are $S'_{r_j+1}(j-1)$ such trees. Let $E$ be the set of elements $x$ that are not in any tree $T_t(j)$ nor in ${\mathfrak{s}}_t(1)$, and let $r = |E| = (n-k-2) - (r_2 + \cdots + r_k)$. By Lemma~\ref{lem:ffseq}, $n \in {\mathfrak{s}}_t(x)$ for all $x \in E$. Then the union of paths $P_t(x)$ for all $x \in E$ form a labeled tree rooted at $n$ with height at most $k$, and there are $S'_{r+1}(k)$ such trees. Thus the number of transformations in this case is $O_n$.
\item $n-1 \not\in {\mathfrak{s}}_t(1)$. Now, for all $j \ge 1$, $1t^j \neq n-1$. Then $t \in \Bbf(n)$. As in the proof of Lemma~\ref{lem:Hn}, the number of transformations in this case is $N_n$.
\end{enumerate}
Altogether we have the desired formula. \qed
\end{proof}
Let $\bff(n) = |\mathbf{B}_{\mathrm{ff}}(n)|$. From Proposition~\ref{prop:ff} and Lemma~\ref{lem:Fn} we have
\begin{proposition}\label{prop:Fncard}
For $n \ge 3$, if $L$ is a factor-free regular language with quotient complexity $n$, then its syntactic complexity $\sigma(L)$ satisfies that $\sigma(L) \le \bff(n)$, where $\bff(n)$ is the cardinality of $\mathbf{B}_{\mathrm{ff}}(n)$ as in Lemma~\ref{lem:Fn}.
\end{proposition}
The tight upper bound on the syntactic complexity of factor-free regular languages is reached by the largest semigroup contained in $\mathbf{B}_{\mathrm{ff}}(n)$. When $2 \le n \le 4$, $\mathbf{B}_{\mathrm{ff}}(n)$ is a semigroup. The languages $L_2 = \varepsilon$, $L_3 = a$ over alphabet $\{a,b\}$, and $L_4 = ab^*a$ have syntactic complexities $1 = \bff(2)$, $2 = \bff(3)$, and $6 = \bff(4)$, respectively. So $\bff(n)$ is a tight upper bound for $n \in \{2,3,4\}$. However, the set $\mathbf{B}_{\mathrm{ff}}(n)$ is not a semigroup for $n \ge 5$, because $s_1 = [2,3,\ldots,n-1,n,n], s_2 = {n-1 \choose n}{2 \choose n-1}{1 \choose n} = [n,n-1,3,\ldots,n-2,n,n] \in \mathbf{B}_{\mathrm{ff}}(n)$ but $s_1s_2 = [n-1,3,\ldots,n-2,n,n,n] \not\in \mathbf{B}_{\mathrm{ff}}(n)$.
Next, we find a large semigroup that can be the syntactic semigroup of a factor-free regular language.
Let $t_0 = {Q \setminus \{1\} \choose n}{1 \choose n-1} = [n-1, n, \ldots, n]$, and let $\Wff(n) = \mathbf{U}^1_n \cup \{t_0\} \cup \mathbf{U}^3_n$. When $2 \le n \le 4$, we have $\Wff(n) = \mathbf{B}_{\mathrm{ff}}(n)$. So we are interested in larger values of $n$ in the rest of this section.
\begin{proposition}\label{prop:Uncard}
For $n \ge 5$, $\Wff(n)$ is a semigroup contained in $\mathbf{B}_{\mathrm{ff}}(n)$ with cardinality
$$\wff(n) = |\Wff(n)| = (n-1)^{n-3} + (n-3)2^{n-3} + 1.$$
\end{proposition}
\begin{proof}
As we have shown in the proof of Proposition~\ref{prop:Rcard}, $\mathbf{U}^1_n$ is a semigroup. For any $t \in \mathbf{U}^1_n \cup \{t_0\}$, since $t_0 \in \mathbf{U}^2_n$, we have $tt_0, t_0t \in \mathbf{U}^1_n$; so $\mathbf{U}^1_n \cup \{t_0\}$ is also a semigroup. We also know that, for any $t_3 \in \mathbf{U}^3_n$ and $t' \in \Wff(n)$, since $\Wff(n) \subseteq \Wbf(n)$, $i(t_3t') = n$ for all $i \neq 1$; so $t_3t' \in \Wff(n)$. If $t' \in \mathbf{U}^1_n \cup \{t_0\}$, then $1t't_3 = n$ and $t't_3 \in \mathbf{U}^1_n$; otherwise, $t' \in \mathbf{U}^3_n$, and $t't_3 = t_2$ or ${Q \choose n} \in \mathbf{U}^1_n$. Hence $\Wff(n)$ is a semigroup.
For any $t \in \mathbf{U}^1_n$, since $1t = n$, we have $t \in \mathbf{B}_{\mathrm{ff}}(n)$. For any $t \in \mathbf{U}^3_n$, $1t \neq n-1$, and $it^2 = n$ for all $i \in \{2,\ldots,n\}$; then $t \in \mathbf{B}_{\mathrm{ff}}(n)$ as well. Clearly $t_0 \in \mathbf{B}_{\mathrm{ff}}(n)$. Hence $\Wff(n)$ is contained in $\mathbf{B}_{\mathrm{ff}}(n)$.
We know that $|\mathbf{U}^1_n| = (n-1)^{n-3}$ and $|\mathbf{U}^3_n| = (n-3)2^{n-3}$. Therefore $|\Wff(n)| = (n-1)^{n-3} + (n-3)2^{n-3} + 1$. \qed
\end{proof}
We now describe a generating set of $\Wff(n)$.
\begin{proposition}\label{prop:Ugen}
For $n \ge 5$, the semigroup $\Wff(n)$ is generated by $\Hff(n) = \{a_1, a_2, a_3, b_1, \ldots, b_{n-3}, c_1, \ldots, c_m\}$, where $m = (n-3)(2^{n-3}-1)$, and
\begin{enumerate}
\item $a_1 = {1 \choose n}{n-1 \choose n}(2,\ldots,n-2)$, $a_2 = {1 \choose n}{n-1 \choose n}(2,3)$, $a_3 = {1 \choose n}{n-1 \choose n}{n-2 \choose 2}$;
\item For $1 \le i \le n-3$, $b_i = {1 \choose n}{n-1 \choose n}{i+1 \choose n-1}$;
\item Each $c_i$ defines a distinct transformation in $\mathbf{U}^3_n$ other than $[j,n,\ldots,n,n]$ for all $j \in \{2,\ldots,n-2\}$.
\end{enumerate}
\end{proposition}
\begin{proof}
We know from the proof of Proposition~\ref{prop:Rgen} that $\mathbf{U}^1_n$ is generated by $\{a_1,a_2,a_3,b_1,\ldots,b_{n-3}\}$. Also, the transformations that are in $\{t_0\} \cup \mathbf{U}^3_n$ but not in $\Hff(n)$ are $t_j = [j,n,\ldots,n,n]$, where $j \in \{2,\ldots,n-1\}$. Each $t_j$ is a composition of $d = {1 \choose 2}{n-1 \choose n}{Q' \choose n-1} \in \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$ and $s_j = {1 \choose n}{n-1 \choose n}{2 \choose j} \in \mathbf{U}^1_n$. Therefore $\langle \Hff(n) \rangle = \Wff(n)$. \qed
\end{proof}
\begin{theorem}\label{thm:ffaut}
For $n \ge 5$, let ${\mathcal A}_n = (Q, \Sigma, \delta, 1, F)$ be the DFA with alphabet $\Sigma = \{a_1, a_2, a_3, b_1, \ldots, b_{n-3}, c_1, \ldots, c_m\}$ of size $(n-3)2^{n-3}+3$, where each letter defines a transformation as in Proposition~\ref{prop:Ugen}, and $F = \{n-1\}$. Then $L = L({\mathcal A}_n)$ has quotient complexity $\kappa(L) = n$, and syntactic complexity $\sigma(L)~=~\wff(n)$. Moreover, $L$ is factor-free.
\end{theorem}
\begin{proof}
Since $\Hff(n) \subseteq \mathbf{G}^{\ge 6}_{\mathrm{bf}}(n)$, the DFA ${\mathcal A}_n$ can be obtained from the DFA ${\mathcal A}'_n$ of Theorem~\ref{thm:bfaut1} by restricting the alphabet. The words used to show that all the states of ${\mathcal A}'$ are reachable and distinct still exist in ${\mathcal A}_n$. Then we have $\kappa(L) = n$. By Proposition~\ref{prop:Ugen}, the syntactic semigroup of $L$ is $\Wff(n)$; so $\sigma(L) = \wff(n)$. By Proposition~\ref{prop:ff}, $L$ is factor-free. \qed
\end{proof}
\begin{conjecture}[Factor-Free Regular Languages]\label{con:ff}
If $L$ is a factor-free regular language with $\kappa(L) = n$, where $n \ge 5$, then $\sigma(L) \le \wff(n)$ and this is a tight upper~bound.
\end{conjecture}
We prove the conjecture for $n = 5$ and $6$.
\begin{proof}
For $n = 5$, $|\mathbf{B}_{\mathrm{ff}}(5)| = 31$, and $|\Wff(5)| = 25$. There are 6 transformations $\tau_1, \ldots, \tau_{6}$ in $\mathbf{B}_{\mathrm{ff}}(5) \setminus \Wff(5)$. For each $\tau_i$, $1 \le i \le 6$, we found a unique $t_i \in \Wff(5)$ such that $\langle t_i, \tau_i \rangle \not\subseteq \mathbf{B}_{\mathrm{ff}}(5)$:
$$\begin{array}{ll}
\tau_1 = [ 2, 3, 4, 5, 5 ], \quad & \quad t_1 = [ 5, 2, 2, 5, 5 ], \\
\tau_2 = [ 2, 3, 5, 5, 5 ], \quad & \quad t_2 = [ 5, 4, 2, 5, 5 ], \\
\tau_3 = [ 2, 5, 3, 5, 5 ], \quad & \quad t_3 = [ 5, 3, 3, 5, 5 ], \\
\tau_4 = [ 3, 2, 5, 5, 5 ], \quad & \quad t_4 = [ 5, 2, 4, 5, 5 ], \\
\tau_5 = [ 3, 4, 2, 5, 5 ], \quad & \quad t_5 = [ 5, 3, 2, 5, 5 ], \\
\tau_6 = [ 3, 5, 2, 5, 5 ], \quad & \quad t_6 = [ 5, 3, 4, 5, 5 ].
\end{array}$$
For each $1 \le i \le 6$, at most one of $t_i$ and $\tau_i$ can appear in the syntactic semigroup $T_L$ of a factor-free regular language $L$. Then $\sigma(L) = |T_L| \le 25$. By Theorem~\ref{thm:ffaut}, this upper bound is tight for $n = 5$.
For $n = 6$, $|\mathbf{B}_{\mathrm{ff}}(6)| = 246$, and $|\Wff(6)| = 150$. There are 96 transformations $\tau_1, \ldots, \tau_{96}$ in $\mathbf{B}_{\mathrm{ff}}(6) \setminus \Wff(6)$. For each $\tau_i$, $1 \le i \le 72$, we enumerated the transformations in $\Wff(6)$ using \emph{GAP} and found a unique $t_i \in \Wff(6)$ such that $\langle t_i, \tau_i \rangle \not\subseteq \mathbf{B}_{\mathrm{ff}}(6)$. Thus $150$ is a tight upper bound for $n = 6$. \qed
\end{proof}
\section{Quotient Complexity of the Reversal of Free Languages}\label{sec:rev}
It has been shown in~\cite{BrYe11} that for certain regular languages with maximal syntactic complexity, the reverse languages have maximal quotient complexity. This is also true for some free languages, as we now show.
In this section we consider \emph{non-deterministic finite automata} (NFA). A NFA ${\mathcal N}$ is a quintuple ${\mathcal N} = (Q, \Sigma, \delta, I, F)$, where $Q$, $\Sigma$, and $F$ are as in a DFA, $\delta : Q \times \Sigma \to 2^Q$ is the non-deterministic transition function, and $I$ is the set of initial states. For any word $w \in \Sigma^*$, the \emph{reverse} of $w$ is defined inductively as follows: $w^R = \varepsilon$ if $w = \varepsilon$, and $w^R = u^Ra$ if $w = au$ for some $a \in \Sigma$ and $u \in \Sigma^*$. The \emph{reverse} of any language $L$ is the language $L^R = \{w^R \mid w \in L\}$. For any finite automaton (DFA or NFA) ${\mathcal M}$, we denote using ${\mathcal M}^R$ the automaton obtained by reversing ${\mathcal M}$ and exchanging the roles of initial states and accepting states, and ${\mathcal M}^D$, the DFA obtained by applying the subset construction to ${\mathcal M}$. Then $L({\mathcal M}^R) = (L({\mathcal M}))^R$, and $L({\mathcal M}^D) = L({\mathcal M})$. To simplify our proofs, we use an observation from~\cite{Brz62} that, for any NFA ${\mathcal N}$ whose states are all reachable, if the automaton ${\mathcal N}^R$ is deterministic, then the DFA ${\mathcal N}^D$ is minimal.
\begin{theorem}\label{thm:pfrev}
The reverse of the prefix-free regular language accepted by
the DFA ${\mathcal A}_n$ of Theorem~\ref{thm:prefix-free} restricted to $\{a,c,d_{n-2}\}$
has $2^{n-2}+1$ quotients, which is the maximum possible for a prefix-free regular language.
\end{theorem}
\begin{proof}\label{proof:pfrev}
Let ${\mathcal B}_n$ be the DFA ${\mathcal A}_n$ restricted to $\{a,c,d_{n-2}\}$. Since $L({\mathcal A}_n)$ is prefix-free, so is $L_n = L({\mathcal B}_n)$. We show that $\kappa(L_n^R) = 2^{n-2}+1$.
Let ${\mathcal N}_n$ be the NFA obtained by removing unreachable states from the NFA ${\mathcal A}_n^R$. (See Fig.~\ref{fig:pfrev} for ${\mathcal N}_6$.) We first prove that the following $2^{n-2}+1$ sets of states of ${\mathcal N}_n$ are reachable:
$\{\{n-1\}\} \cup \{S \mid S \subseteq \{1,\ldots,n-2\}~\}.$
\begin{figure}[hbt]
\begin{center}
\input{rev6.eepic}
\end{center}
\caption{NFA ${\mathcal N}_6$ of $L_6^R$ with quotient complexity $\kappa(L_6^R) = 17$; empty state omitted.}
\label{fig:pfrev}
\end{figure}
The singleton set $\{n-1\}$ of initial states of ${\mathcal N}_n$ is reached by $\varepsilon$. From $\{n-1\}$ we reach the empty set by $a$.
The set $\{n-2\}$ is reached by $d_{n-2}$ from $\{n-1\}$, and from here, $\{1\}$ is reached by $a^{n-3}$. From any set $\{1,2,\ldots,i\}$, where $1\le i<n-2$, we reach
$\{1,2,\ldots,i, i+1\}$ by $ca^{n-3}$. Thus we reach $\{1,2,\ldots,n-2\}$ from $\{1\}$ by
$(ca^{n-3})^{n-3}$.
Now assume that any set $S$ of cardinality $l\le n-2$ can be reached; then we can get a set of cardinality $l-1$ by deleting an element~$j$ from $S$ by applying
$a^jd_{n-2}a^{n-2-j}$. Hence all the subsets of $\{1,2,\ldots,n-2\}$ can be reached.
The automaton ${\mathcal N}_n^R$ is a subset of ${\mathcal A}_n$, and it is deterministic. Then ${\mathcal N}_n^D$ is minimal. Hence $\kappa(L_n^R) = 2^{n-2}+1$, which is the maximal quotient complexity of reversal of prefix-free languages as shown in~\cite{HSW09}. \qed
\end{proof}
\medskip
It is interesting that, for suffix-, bifix-, and factor-free regular languages, although we don't have tight upper bounds on their syntactic complexities, some languages in these classes with large syntactic complexities have their reverse languages reaching the upper bounds on the quotient complexities for the reversal operation.
\begin{theorem}\label{thm:sfrev}
The reverse of the suffix-free regular language accepted by the DFA ${\mathcal A}_n'$ of Theorem~\ref{thm:wsfaut} restricted to $\{a_1,a_2,a_3,c\}$ has $2^{n-2}+1$ quotients, which is the maximum possible for a suffix-free regular language.
\end{theorem}
\begin{proof}
Let ${\mathcal C}_n$ be the DFA ${\mathcal A}_n'$ restricted to the alphabet $\{a_1,a_2,a_3,c\}$. Since $L({\mathcal A}_n')$ is suffix-free, so is $L_n' = L({\mathcal C}_n)$. Let ${\mathcal N}_n'$ be the NFA obtained from ${\mathcal C}_n^R$ by removing unreachable states. Figure~\ref{fig:sfrev} shows the NFA ${\mathcal N}_6'$.
\begin{figure}[hbt]
\begin{center}
\input{sfrev.eepic}
\end{center}
\caption{NFA ${\mathcal N}_6'$ of $L_6'^R$ with quotient complexity $\kappa(L_6'^R) = 17$; empty state omitted.}
\label{fig:sfrev}
\end{figure}
Apply the subset construction to ${\mathcal N}_n'$, we get a DFA ${\mathcal N}_n'^D$. Its initial state is a singleton set $\{2\}$. From the initial state, we can reach state $\{2,3,\ldots,i\}$ by $(a_3a_1^{n-3})^{i-2}$, where $3 \le i \le n-1$. Then the state $\{2,3,\ldots,n-1\}$ is reached from $\{2\}$ by $(a_3a_1^{n-3})^{n-3}$. Assume that any set $S$ of cardinality $l$ can be reached, where $2 \le l \le n-2$. If $j \in S$, then we can reach $S' = S \setminus \{j\}$ from $S$ by $a_1^{j-1}a_3a_1^{n-j-1}$. So all the nonempty subsets of $\{2,3,\ldots,n-1\}$ can be reached. We can also reach the singleton set $\{1\}$ from $\{2\}$ by $c$, and, from there, the empty state by $c$ again. Hence ${\mathcal N}_n'^D$ has $2^{n-2}+1$ reachable states.
Since the automaton ${\mathcal N}_n'^R$, the reverse of ${\mathcal N}_n'$, is a subset of ${\mathcal C}_n$, it is deterministic; hence ${\mathcal N}_n'^D$ is minimal. Then the quotient complexity of $L_n'^R$ is $2^{n-2}+1$, which meets the upper bound for reversal of suffix-free regular languages~\cite{HS09}. \qed
\end{proof}
\begin{theorem}\label{thm:ffrev}
The reverse of the factor-free regular language accepted by the DFA ${\mathcal A}_n$ of Theorem~\ref{thm:ffaut} restricted to the alphabet $\{a_1,a_2,a_3,c\}$, where $c = [2,n-1,n,\ldots,n,n] \in \Hff(n)$, has $2^{n-3}+2$ quotients, which is the maximum possible for a bifix- or factor-free regular language.
\end{theorem}
\begin{proof}
Let ${\mathcal D}_n$ be the DFA ${\mathcal A}_n$ restricted to the alphabet $\{a_1,a_2,a_3,c\}$; then $L_n'' = L({\mathcal D}_n)$ is factor-free. Let ${\mathcal N}_n''$ be the NFA obtained from ${\mathcal D}_n^R$ by removing unreachable states. An example of ${\mathcal N}_n''$ is shown in Figure~\ref{fig:ffrev}.
\begin{figure}[hbt]
\begin{center}
\input{ffrev.eepic}
\end{center}
\caption{NFA ${\mathcal N}_7''$ of $L_7''^R$ with quotient complexity $\kappa(L_7''^R) = 18$; empty state omitted.}
\label{fig:ffrev}
\end{figure}
Note that ${\mathcal N}_n''$ can be obtained from the NFA ${\mathcal N}_{n-1}'$ in Theorem~\ref{thm:sfrev} by adding a new state $n-1$, which is the only initial state in ${\mathcal N}_n''$, and the transition from $\{n-1\}$ to $\{2\}$ under input $c$. We know that all non-empty subsets of $\{2,3,\ldots,n-2\}$ are reachable from $\{2\}$. The accepting state $\{1\}$ is also reachable from $\{2\}$. From the initial state $n-1$, we reach the empty state under input $a_1$. Then ${\mathcal N}_n''^D$ has $2^{n-3}+2$ reachable states.
Since ${\mathcal N}_n''^R$ is a subset of ${\mathcal D}_n$ and it is deterministic, the DFA ${\mathcal N}_n''^D$ is minimal. Therefore $\kappa(L_n''^R) = 2^{n-3}+2$, and it reaches the upper bound for reversal of both bifix- and factor-free regular languages with quotient complexity $n$~\cite{BJLS11}. \qed
\end{proof}
\section{Conclusions}\label{sec:cl}
Our results are summarized in Tables~\ref{tab:Summary1} and~\ref{tab:Summary2}. Each cell of Table~\ref{tab:Summary1} shows the syntactic complexity bounds of prefix- and suffix-free regular languages, in that order, with a particular alphabet size. Table~\ref{tab:Summary2} is structured similarly for bifix- and factor-free regular languages. The figures in bold type are tight bounds verified by {\it GAP}. To compute the bounds for suffix-, bifix-, and factor-free languages, we enumerated semigroups generated by elements of $\Bsf(n)$, $\Bbf(n)$, and $\mathbf{B}_{\mathrm{ff}}(n)$ that are contained in $\Bsf(n)$, $\Bbf(n)$, and $\mathbf{B}_{\mathrm{ff}}(n)$, respectively, and recorded the largest ones. By Propositions~\ref{prop:sf},~\ref{prop:bf},~\ref{prop:ff}, we obtained the desired bounds from the enumeration. The asterisk $\ast$ indicates that the bound is already tight for a smaller alphabet. In Table~\ref{tab:Summary1}, the last four rows include the tight upper bound $n^{n-2}$ for prefix-free languages, $\vsf(n)$, which is a tight upper bound for $2 \le n \le 5$ for suffix-free languages, conjectured upper bound ${\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n)$ for suffix-free languages, and a weaker upper bound $\bsf(n)$ for suffix-free languages. In Table~\ref{tab:Summary2}, the last four rows include $\vbf(n)$, which is a tight upper bound for bifix-free languages for $2 \le n \le 5$, conjectured upper bounds $\wbf(n)$ for bifix-free languages and $\wff(n)$ for factor-free languages, and weaker upper bounds $\bbf(n)$ for bifix-free languages and $\bff(n)$ for factor-free languages.
\vspace{-.4cm}
\begin{table}[H]
\caption{Syntactic complexities of prefix- and suffix-free regular languages.}
\label{tab:Summary1}
\begin{center}
$
\begin{array}{|c||c|c|c|c|c|}
\hline
\ \ \ \ &\ \ n=2 \ \ &\ \ n=3 \ \ & \ \ n=4 \ \
& \ n=5 \ & \ n=6 \ \\
\hline \hline
|\Sigma|=1
& {\bf 1} & {\bf 2} & {\bf 3} & {\bf 4}
& {\bf 5} \\
\hline
|\Sigma|=2
& \ast & {\bf 3}/{\bf 3} & {\bf 11}/{\bf 11}
& {\bf 49}/{\bf 49} & ? \\
\hline
|\Sigma|=3
& \ast & \ast & {\bf 14}/{\bf 13}
& {\bf 95}/{\bf 61} & ? \\
\hline
|\Sigma|=4
& \ast & \ast & \hspace{-.2cm}{\bf 16}/\ast
& \hspace{-.1cm}{\bf 110}/{\bf 67} & ? \\
\hline
|\Sigma|=5
& \ast & \ast & \ast
& \hspace{-.1cm}{\bf 119}/{\bf 73} & ? \\
\hline
|\Sigma|=6
& \ast & \ast & \ast
& \hspace{-.2cm}{\bf 125}/~\ast & ~~?~/501 \\
\hline
|\Sigma|=7
& \ast & \ast & \ast
& \ast & \hspace{-.5cm}~{\bf 1296}/~?~ \\
\hline
|\Sigma|=8
& \ast & \ast & \ast
& \ast & ~~\ast~/{\bf 629} \\
\hline
\cdots
& & &
& & \\
\hline
~n^{n-2}~
& \hspace{-.3cm}{\bf 1} & \hspace{-.3cm}{\bf 3} & \hspace{-.5cm}{\bf 16}
& \hspace{-.6cm}{\bf 125} & \hspace{-.9cm}{\bf 1296} \\
\hline
~\vsf(n)~
&\hspace{.3cm}{\bf 1} &\hspace{.3cm}{\bf 3} &\hspace{.5cm}{\bf 13}
&\hspace{.6cm}{\bf 73} &\hspace{.8cm}501\\
\hline
~{\mathrm{w}^{\ge 6}_{\mathrm{sf}}}(n)~
&\hspace{.3cm}{\bf 1} &\hspace{.3cm}{\bf 3} &\hspace{.5cm}11
&\hspace{.6cm}67 &\hspace{.8cm}{\bf 629}\\
\hline
\bsf(n)
&\hspace{.3cm}{\bf 1} &\hspace{.3cm}{\bf 3} &\hspace{.5cm}15
&\hspace{.6cm}115 &\hspace{.8cm}1169\\
\hline
\end{array}
$
\end{center}
\label{table1}
\end{table}
\vspace{-1.4cm}
\begin{table}[H]
\caption{Syntactic complexities of bifix- and factor-free regular languages.}
\label{tab:Summary2}
\begin{center}
$
\begin{array}{|c||c|c|c|c|c|}
\hline
\ \ \ \ &\ \ n = 2 \ \ & \ \ n=3 \ \ & \ \ n=4 \ \
& \ n=5 \ & \ n=6 \ \\
\hline \hline
|\Sigma|=1
& {\bf 1} & {\bf 2} & {\bf 3}
& {\bf 4} & {\bf 5} \\
\hline
|\Sigma|=2
& \ast & \ast & {\bf 7}/{\bf 6}
& {\bf 20}/{\bf 12} & ? \\
\hline
|\Sigma|=3
& \ast & \ast & \ast
& {\bf 31}/{\bf 16} & ? \\
\hline
|\Sigma|=4
& \ast & \ast & \ast
& {\bf 32}/{\bf 19} & ? \\
\hline
|\Sigma|=5
& \ast & \ast & \ast
& {\bf 33}/{\bf 20} & ? \\
\hline
|\Sigma|=6
& \ast & \ast & \ast
& {\bf 34}/~?~ & ? \\
\hline
\cdots
& & &
& & \\
\hline
~\vbf(n)~
&\hspace{-.35cm}{\bf 1} &\hspace{-.35cm}{\bf 2} &\hspace{-.35cm}{\bf 7}
&\hspace{-.6cm}{\bf 34} &\hspace{-.6cm}209 \\
\hline
~\wbf(n)~
&\hspace{-.35cm}{\bf 1} &\hspace{-.35cm}{\bf 2} &\hspace{-.35cm}{\bf 7}
&\hspace{-.6cm}33 &\hspace{-.6cm}{\bf 213} \\
\hline
~\wff(n)~
&\hspace{.35cm}{\bf 1} &\hspace{.35cm}{\bf 2} &\hspace{.35cm}{\bf 6}
&\hspace{.5cm}{\bf 25} &\hspace{.65cm}{\bf 150} \\
\hline
\bbf(n)/\bff(n)
&{\bf 1}/{\bf 1} &{\bf 2}/{\bf 2} &{\bf 7}/{\bf 6}
&41/31 &339/246 \\
\hline
\end{array}
$
\end{center}
\label{table2}
\end{table}
|
\section{Introduction}
\setcounter{equation}{0}\setcounter{theorem}{0}
Let $\g$ be a finite-dimensional complex semisimple Lie algebra and
$U(\g)$ its universal enveloping algebra. The study of the
$\g$-module structure of its exterior algebra $\w\g$ has a long
history. Although this module structure is still not fully
understood, Kostant
have done a lot of important work on it, see for example \cite{k1} and \cite{k2}.
Let $Cas\in U(\g)$ be the Casimir element with respect to the
Killing form. Let $m_i$ be the maximal eigenvalue of $Cas$ on
$\wedge ^{i} \g$ and $M_i$ be the corresponding eigenspace. Let $p$
be the maximal dimension of commutative subalgebras of $\g$. In
\cite{k1} it is proved that $m_i\le i$ for any $i$ and $m_i=i$ for
$0\le i\le p$, and if $m_i=i$ then $M_i$ is a multiplicity-free
$\g$-module whose highest weight vectors corresponding to
$i$-dimensional abelian ideals of $\b$. The integer $p$ for all the
simple Lie algebras was determined by Malcev, and Suter
gave a uniform formula for $p$ in \cite{s}.
Fix a Cartan subalgebra $\h$ of $\g$ and a set $\ddt^+$ of positive
roots. Let $\rho\in \h^{*}$ be one half the sum of all the positive
roots. For any $\ld\in\h^{*}$, let $V_{\ld}$ denote the irreducible
representation of $\g$ with highest weight $\ld$.
In this paper we will prove the following result, which extends some theorems of Kostant. Let $n=dim \ \g$ and $r$ be the number of positive
roots.
\begin{theorem} [Theorem \ref{e}]
One has $m_i\le n/3$ for $i=0,1,\cdots,n$, and $m_i=n/3$ if and only if
$i=r,r+1,\cdots,r+l$. For $s=0,1,\cdots,l$, $M_{r+s}=\left(
\begin{array}{c}
l \\
s \\
\end{array}
\right) V_{2\rho}$.
For $0\le i< r$ one has $m_i< m_{i+1}$. For $1\le i\le r$, $M_i$ is a multiplicity-free $\g$-module, whose highest
weight vectors corresponding to certain ad-nilpotent ideals of $\b$. In fact $\oplus_{i=0}^r M_i$ is also a
multiplicity-free $\g$-module.
\end{theorem}
This result relates $M_i$ to ad-nilpotent ideals of $\b$, which are
classified in \cite{pp}. But it will be complicated to
determine those ad-nilpotent ideals of $\b$ corresponding to the
highest weight vectors of $M_i$.
To prove this theorem, we need the following interesting result.
\begin{prop}[Proposition \ref{d}]
Let $k\in \mathbb{Z}^+$. The set of weights of $V_{k\rho}$ (whose
dimension is $(k+1)^r$) is $$\{\sum_{i=1}^r c_i\al_i|\al_i\in\ddt^+,
c_i=-k/2,-k/2+1,\cdots,k/2-1,k/2\}.$$
\end{prop}
\section{Weights of a representation with highest weight a multiple of $\rho$}
\setcounter{equation}{0}\setcounter{theorem}{0}
Let $\g$ be a finite-dimensional complex semisimple Lie algebra. Fix
a Cartan subalgebra $\h$ of $\g$ and a Borel subalgebra $\b$ of $\g$
containing $\h$. Let $\ddt$ be the set of roots of $\g$ with respect
to $\h$ and $\ddt^+$ be the set of positive roots whose
corresponding root spaces lie in $\b$. Let $W$ be the Weyl group.
Let $\n=[\b,\b]$. Then $\b=\h\oplus\n$. An ideal of $\b$ contained
in $\n$ is called an ad-nilpotent ideal of $\b$, as it consists of
ad-nilpotent elements. Let $\Gamma\st \h^{*}$ be the lattice of
$\g$-integral linear forms on $\h$ and $\Lambda\st\Gamma$ the subset
of dominant integral linear forms. Let $(,)$ be the bilinear form on
$\h^{*}$ induced by the Killing form. Let $l=dim\ \h$, $r=|\ddt^+|$
and $n=l+2r=dim\ \g$. Assume $\ddt^+=\{\al_1,\al_2,\cdots,\al_r\}.$
For any $\ld\in \Lambda$,
let $\pi_\ld:\g\rt End(V_\ld)$ be the irreducible representation of
$\g$ with highest weight $\ld$, and $\Ga(V_\ld)$ be the set of
weights, with multiplicities. Any $\ga\in\Ga(V_\ld)$ will appear $k$
times if the dimension of the $\ga$-weight space is $k$. For example
$\Ga(\g)=\ddt\cup\{0,\cdots,0\}$ ($l$ times). If $U\st V_\ld$ is an
$\h$-invariant
subspace then we will also use $\Ga(U)$ to denote the the set of weights of $U$ with multiplicities and define $$<U>=\sum_{\ga\in\Ga(U)} \ga.$$
For any $S\st \Ga(V_\ld)$, we also define $<S>=\sum_{\ga\in S} \ga.$
Let $\rho\in
\h^{*}$ be one half the sum of all the positive roots. For any $k\in
\mathbb{Z}^+$, the representation $V_{k\rho}$ of $\g$ has dimension
$(k+1)^r$ by Weyl's dimension formula. The following result describes the set of weights of
$V_{k\rho}$, which is well-known if $k=1$ (see e.g. \cite{k0}).
\begin{prop}\lb{d}
The set of weights of $V_{k\rho}$ is $$\Ga(V_{k\rho})=\{\sum_{i=1}^r
c_i\al_i|\al_i\in\ddt^+, c_i=-k/2,-k/2+1,\cdots,k/2-1,k/2\},$$ or
equivalently,
$$\Gamma(V_{k\rho})=\{k\rho-\sum_{i=1}^r c_i\al_i|\al_i\in\ddt^+,
c_i=0,1,\cdots,k.\}.$$
\end{prop}
\bp By Weyl's denominator formula
$$\prod_{i=1}^r (e^{\frac{k+1}{2}\al_i}-e^{-\frac{k+1}{2}\al_i})=\sum_{w\in W} sgn(w)
e^{w((k+1)\rho)}.$$ Then for $c_i=-k/2,-k/2+1,\cdots,k/2-1,k/2$ with
$i=1,\cdots,r$,
\bee
\begin{split}
\sum_{c_1,\cdots,c_r} e^{\sum_{i=1}^r
c_i\al_i}&=\prod_{i=1}^r(e^{(-\frac{k}{2})\al_i}+e^{(-\frac{k}{2}+1)\al_i}+\cdots+e^{(\frac{k}{2}-1)\al_i}+e^{(\frac{k}{2})\al_i})\\
&=\prod_{i=1}^r\frac{e^{\frac{k+1}{2}\al_i}-e^{-\frac{k+1}{2}\al_i}}{e^{\frac{1}{2}\al_i}-e^{-\frac{1}{2}\al_i}}\\&=\frac{\sum_{w\in
W} \ sgn(w) e^{w((k+1)\rho)}}{\prod_{i=1}^r
(e^{\frac{1}{2}\al_i}-e^{-\frac{1}{2}\al_i})}\\&=char(V_{k\rho}).
\end{split}
\eee \ep
Let $Cas\in U(\g)$ be the Casimir element corresponding
to the Killing form. For any $\ld\in\Ga$, define
$$Cas(\ld)=(\ld+\rho,\ld+\rho)-(\rho,\rho).$$ The following result is well-known.
\begin{lem}\lb{c}
If $\ld\in\Lambda$ then $Cas(\ld)$ is the scalar value
taken by $Cas$ on $V_\ld$. For any $\mu\in \Ga(\ld)$ one has $Cas(\mu)\le Cas(\ld)$ and $Cas(\mu)<Cas(\ld)$ if $\mu\neq\ld$.
\end{lem}
\section{Maximal eigenvalues of a Casimir operator and the corresponding eigenspaces}
\setcounter{equation}{0}\setcounter{theorem}{0}
Let $\wedge \g$ be the exterior algebra of $\g$. Then $\g$ acts on
$\wedge \g$ naturally. Let $m_i$ be the maximal eigenvalue of $Cas$
on $\wedge ^{i} \g$ and $M_i$ be the corresponding eigenspace.
One knows that $\wedge ^{i}\g$ is isomorphic to $\w
^{n-i}\g$ as $\g$-modules for each $i$, so one has
$$m_i=m_{n-i}$$ and $$M_i\cong M_{n-i}.$$ Let $p$ be the maximal dimension of commutative
subalgebras of $\g$. Kostant showed that $m_i\le i$ and $m_i=i$ for
$0\le i\le p$, and if $m_i=i$ then $M_i$ is spanned by $\wedge ^{k}
\a$, where $\a$ runs through $k$-dimensional commutative subalgebras
of $\g$.
A nonzero vector $w\in\w\g$ is called decomposable if
$w=z_1\w z_2\w\cdots\w z_k$ for some
positive integer $k$, where $z_i\in \g$. In this case let
$\a(w)$ be the respective $k$-dimensional subspace spanned by
$z_1,z_2,\cdots,z_k$.
\begin{theorem} [Proposition 6 and Theorem 7 of \cite{k1}]\lb{b}
(1) Let $$w=z_1\w z_2\w\cdots\w z_k\in\w^k\g$$ be a decomposable
vector. Then $w$ is a highest weight vector if and only if $\a(w)$
is $\b$-normal, i.e., $[\b,\a(w)]\st\a(w)$. In this case the highest
weight of the simple $\g$-module generated by $w$ is $<\a(w)>$.
Thus there is a one-to-one correspondence between all the decomposably-generated simple $\g$-submodules of $\w^k\g$ and all the
$k$-dimensional $\b$-normal subspaces of $\g$.
(2) Let $\a_1,\a_2$ be any two ideals of $\b$ lying in $\n$. Then
$<\a_1>=<\a_2>$ if and only if $\a_1=\a_2$. Thus, if
$V_1\st\w^k\g,V_2\st\w^j\g$ are two decomposably-generated simple
$\g$-submodules which corresponds to ideals of $\b$ lying in $\n$,
then $V_1$ is equivalent to $V_2$ if and only if $V_1=V_2$.
\end{theorem}
\begin{theorem} \lb{e} (1) One has $$m_i=max\{||\rho+\ga_1+\cdots+\ga_i||^2-||\rho||^2\ |\{\ga_t|t=1,\cdots,i\}\st\Ga(\g)\}$$ for any $i$.
(2)One has $m_i\le n/3$ for $i=0,1,\cdots,n$, and $m_i=n/3$ if and
only if $i=r,r+1,\cdots,r+l$. For $s=0,1,\cdots,l$, $M_{r+s}=\left(
\begin{array}{c}
l \\
s \\
\end{array}
\right) V_{2\rho}$.
(3)For $0\le k< r$ one has $m_k< m_{k+1}$. For $1\le k\le r$, $M_k$ is a multiplicity-free $\g$-module, whose highest
weight vectors corresponding to those $k$-dimensional ad-nilpotent ideals $\a$ of $\b$ such that $Cas(<\a>)=m_k$. In fact $\oplus_{k=0}^r M_k$ is also a
multiplicity-free $\g$-module.
\end{theorem}
\bp (1) For $j=1,\cdots,r,$ let $x_j$ (resp. $y_j$) be a weight
vector corresponding to $\al_j$ (resp. $-\al_j$). Let
$\{h_1,\cdots,h_l\}$ be a basis of $\h$. Then
$$A=\{x_1,\cdots,x_r,y_1,\cdots,y_r,h_1,\cdots,h_l\}$$ is a basis of
$\g$ consisting of weight vectors. Then $$B_i=\{a_1\wedge a_2\wedge
\cdots\wedge a_i| a_j\in A\}$$ is a basis of $\w^{i}\g$ consisting
of weight vectors. Let $$C_i=\{v\in B_i\ | Cas(<\a(v)>)=m_i\}.$$
Then by Corollary 2.1 of \cite{k1} $M_i$ is the direct sum of simple
$\g$-modules with highest weight vectors $v\in C_i$. It is clear
that $$Cas(<\a(v)>)=||\rho+\ga_1+\cdots+\ga_i||^2-||\rho||^2$$ if
the weight of $a_j$ is $\ga_j$, thus (1) follows.
(2)For any $S=\{\ga_j|j=1,\cdots,i\}\st\Ga(\g)$, $<S>$ is a weight
of $\pi_{2\rho}$ by Proposition \ref{d}. Thus by Lemma \ref{c}
$Cas(<S>)\le Cas(2\rho)=8||\rho||^2=n/3$, as $||\rho||^2=n/24$. So
$m_i=n/3$ if and only if there exists $S\st\Ga(\g)$ such that
$|S|=i$ and $<S>=2\rho$. Then $S$ must be of the form
$\{x_1,\cdots,x_r,h_{j_1},\cdots,h_{j_s}\}$ and thus $r\le i\le
r+l$. For $0\le s\le l$, it is clear $$C_{r+s}=\{x_1\w\cdots\w x_r\w
h_{j_1}\w\cdots\w h_{j_s}|1\le j_1<j_2<\cdots<j_s\le l \},$$ thus
$M_{r+s}=\left(
\begin{array}{c}
l \\
s \\
\end{array}
\right) V_{2\rho}$.
(3)We first show $m_{k+1}>m_{k}$ for $0\le k<r$, which clearly holds
in the case $k=0$. Assume $1\le k<r$. Let $v=a_1\w\cdots\w a_k\in
C_k$. Then $v$ is a highest weight vector of $M_k$, whose weight is
$<S>$ with $S=\Ga(\a(v))$. Then $Cas(<S>)=m_k$, and
$[\b,\a(v)]\st\a(v)$ by Theorem \ref{b} (1). Recall that for
$\ga=\sum_{i=1}^l k_i \ga_i\in\ddt^+$ where $\{\ga_i|i=1,\cdots,l\}$
is the set of simple roots, its height is defined as $\sum_{i=1}^l
k_i$. Choose a positive root $\al$ in $\ddt^+\setminus S$ (which is
nonempty as $k<r$) with largest height. Set $T=S\cup \{\al\}$. Let
$a\in A$ be the $\al$-weight vector and let $u=v\w a\in B_{k+1}$. By
the choice of $\al$ it is clear that $[\b,\a(u)]\st\a(u)$, thus $u$
is also a highest weight vector, whose weight is $<T>=<S>+\al$. As
$$(<T>,\al)=(<S>,\al)+(\al,\al)>0,$$ $<S>\in \Ga(V_\ld)$ with $\ld={<T>}$. Then
$$m_{k+1}\geq Cas(<T>)>Cas(<S>)=m_k.$$
Now assume $1\le k\le r$. Let $v=a_1\w\cdots\w a_k\in C_k$ , and let
$S=\Ga(\a(v))$. We will show $S\st\ddt^+$. If not, let
$S^{'}=S\setminus(S\cap(-S))$. Then $<S^{'}>=<S>$ and $|S^{'}|=t<k$.
Thus $m_k=Cas(S)=Cas(S^{'})\le m_t$, which contradicts to the
previous result. Thus for $1\le k\le r$ one always has $S\st\ddt^+$.
Any $v\in C_k$ is a highest weight vector, so $[\b,\a(v)]\st \a(v)$.
And if $1\le k\le r$ we have just showed $\Ga(\a(v))\st \ddt^+$.
Thus $\a(v)$ is an ad-nilpotent ideal of $\b$. Let $\ld(v)=<\a(v)>$.
Then
$$M_k=\oplus_{v\in C_k}V_{\ld(v)}.$$ By Theorem \ref{b} (2), if
$v_1,v_2\in C_k$ with $v_1\neq v_2$, then $\a(v_1)\neq\a(v_2)$ and
$\ld(v_1)\neq \ld(v_2)$. Thus $M_k$ is a multiplicity-free
$\g$-module, whose highest weight vectors corresponding to the
ad-nilpotent ideals $\a$ of $\b$ such that $Cas(<\a>)=m_k$. By
Theorem \ref{b} (2) one can further get that $\oplus_{k=0}^r M_k$ is
also a multiplicity-free $\g$-module.
\ep
\begin{rem}
Considering the isomorphism of $\g$-modules $\w^k\g$ and
$\w^{n-k}\g$, $\w^k\g$ is multiplicity-free for $0\le k\le r$ and
$n-r\le k\le n$. For $r\le k\le r+l$ ($r+l=n-r$), we have showed
that $M_k$ is primary of type $\pi_{2\rho}$. As $\g$-modules one has
$\w\g=2^l V_\rho\otimes V_\rho$ (see \cite{k2}), so $\w\g$ contains
exactly $2^l$ copies of $V_{2\rho}$, which is just $\oplus_{s=0}^l
M_{r+s}$.
\end{rem}
|
\section{Introduction}
The one-loop effective action in a background field is an important quantity in quantum field theory \cite{Jackiw:1974cv,Iliopoulos:1974ur}. In gauge theories, the one-loop effective action has been calculated exactly and analytically only for certain very special gauge field backgrounds, such as those with constant (or, in non-Abelian theories, covariantly constant) field strength, based on the seminal work of Heisenberg and Euler \cite{Heisenberg:1935qt}, and Schwinger \cite{Schwinger:1951nm}; and for some very special one-dimensional cases where the field is inhomogeneous \cite{Dunne:2004nc}. The goal of this paper is to extend further the class of background fields for which we can compute the renormalized one-loop effective action. We are particularly interested in studying the small fermion mass limit, motivated by its importance for chiral physics, and also by the fact that while the large mass limit is well understood \cite{Novikov:1983gd}, much less is known about the small mass limit for general gauge backgrounds.
Recently, a new approach has been developed for computing exactly, but numerically, the gauge theory effective action in a class of background fields that permit a separation of variables reducing to a set of one-dimensional operators. This has been explored in most detail for radially separable backgrounds, where the method has been called the "partial-wave cutoff method"
\cite{Dunne:2004sx,Dunne:2005te,Dunne:2006ac,Dunne:2007mt,Hur:2008yg}. The first application of this method was to the full mass dependence of the quark determinant in an instanton background, yielding a smooth interpolation between the known large and small mass extreme limits \cite{Dunne:2004sx}. Subsequent papers have concentrated on scalar theories. Indeed, even
the instanton background computation made use of a special property of self-dual backgrounds that implies that the spinor effective action can be expressed directly in terms of the scalar effective action \cite{'tHooft:1976fv,Jackiw:1977pu,Brown:1977bj,Carlitz:1978xu}. Here, in this paper, we present the spinor approach without assuming self-duality of the background gauge field.
The basic idea of the "partial-wave cutoff method" is simple: the one-loop effective action requires the logarithm of the determinant of an operator, and there is a simple method [known as the Gel'fand-Yaglom method \cite{Gelfand:1959nq}] for computing the determinant of an {\it ordinary} differential operator, without computing its eigenvalues. For a {\it partial} differential operator, if the problem is separable down to a set of {\it ordinary} differential operators, we can formally sum over the Gel'fand-Yaglom results for each term in the separation sum. The technical difficulty is that this sum is naively divergent, and this must be addressed by a suitable regularization and renormalization. This has been resolved in previous publications for gauge theories with scalars and for self-interacting scalar theories. Here we show how this works for spinor theories with background gauge fields that are radially separable, and not necessarily self-dual.
While this class of radially separable backgrounds is large, and includes important special cases such as instantons, monopoles and vortices, there are of course many other physically important background fields that are not separable in this way. In these cases we must resort to approximation methods in order to compute the effective action.
In order to investigate the region of validity of these approximations, we compare our exact numerical results with two such approximations, the large mass expansion and the derivative expansion, and show that for spinor theories a new feature arises when evaluating the derivative expansion approximation for light fermions.
\section{Partial-wave decomposition of the Dirac operator}
We begin with a chiral decomposition of the effective action. The Euclidean one-loop effective action in spinor QED is the logarithm of the determinant of the corresponding Dirac operator:
\begin{eqnarray}
\Gamma[A] = -\ln \, \det\left( \displaystyle{\not} D+m \right)
=-\frac{1}{2}\ln\,\det\left(-\displaystyle{\not} D^2 +m^2\right)
\label{oneloop}
\end{eqnarray}
Here $\displaystyle{\not} D
= \gamma_\mu( \partial_\mu + i A_\mu (x) )
$ is the Dirac operator in Euclidean 4-dimensional spacetime, and $A_\mu(x)$ is the classical background gauge field. We will use the following standard representation for the $4\times 4$ Dirac matrices in Euclidean spacetime \cite{Jackiw:1977pu}:
\begin{equation}
\gamma_\mu = \begin{pmatrix}
& 0 & \alpha_\mu & \\
& \bar{\alpha}_\mu & 0 & \end{pmatrix}
\end{equation}
where
\begin{eqnarray}
\alpha_\mu =(-i \vec{\sigma}, 1) \quad , \quad \bar{\alpha}_\mu =
(\alpha_\mu)^\dagger = (i \vec{\sigma}, 1)
\end{eqnarray}
and the $\sigma_i$ are the $2\times 2$ Pauli matrices. Using this Dirac algebra representation we see clearly the chiral decomposition:
\begin{eqnarray}
-\displaystyle{\not} D^2 +m^2=
\begin{pmatrix}m^2+D D^\dagger &0\cr
0& m^2+D^\dagger D
\end{pmatrix}
\end{eqnarray}
where we have defined
\begin{eqnarray}
D\equiv \alpha_\mu D_\mu\qquad , \qquad D^\dagger \equiv - \bar\alpha_\mu D_\mu
\label{d}
\end{eqnarray}
Thus, we write
\begin{eqnarray}
\Gamma[A] &=&-\frac{1}{2}\ln\,\det\left( m^2+D D^\dagger \right) -\frac{1}{2}\ln\,\det\left( m^2+D^\dagger D \right)\nonumber\\
&\equiv& \Gamma^{(+)}[A]+\Gamma^{(-)}[A]
\label{sum}
\end{eqnarray}
For later use, we recall the familiar properties of the $\alpha_\mu$ matrices:
\begin{eqnarray}
m^2+D D^\dagger &=& m^2-D_\mu^2+\frac{1}{2}F_{\mu\nu} \bar{\eta}^a_{\mu\nu}\,\sigma_a
\label{dddag}\\
m^2+D^\dagger D &=& m^2-D_\mu^2+\frac{1}{2} F_{\mu\nu} \eta^a_{\mu\nu}\,\sigma_a
\label{ddagd}
\end{eqnarray}
where $\eta^a_{\mu\nu}$ and $\bar\eta^a_{\mu\nu}$ are the 't Hooft tensors \cite{'tHooft:1976fv,Jackiw:1977pu}.
Now we make the following simple observation, that we can write the renormalized effective action as
\begin{eqnarray}
\Gamma_{\rm ren}[A] &=& 2\Gamma^{(\pm)}_{\rm ren}[A]
\mp\left(\Gamma^{(+)}_{\rm ren}[A]-\Gamma^{(-)}_{\rm ren}[A]\right)
\label{choice}
\end{eqnarray}
Furthermore, we know that the difference of the renormalized effective action for the two chiralities takes a special form, related to the chiral anomaly:
\begin{eqnarray}
\Delta\Gamma_{\rm ren}[A] &\equiv& \left(\Gamma^{(+)}_{\rm ren}[A]-\Gamma^{(-)}_{\rm ren}[A]\right)\nonumber\\
&=&\frac{1}{2}\frac{1}{(4\pi)^2}\ln \left(\frac{m^2}{\mu^2}\right)\int d^4 x\, F_{\mu\nu} F^*_{\mu\nu}
\label{difference}
\end{eqnarray}
Thus, to evaluate the spinor effective action it is sufficient to evaluate the effective action for just one of the chiralities. That is, we can compute either $\Gamma^{(+)}$ or $\Gamma^{(-)}$, but we do not need to compute both \cite{Hur:2010bd}.
This is a useful observation because there exist background fields for which the computation is significantly easier for one chirality than for the other. For example, if the background field is self-dual then $F_{\mu\nu} \bar{\eta}^a_{\mu\nu}=0$, since $\bar{\eta}^a_{\mu\nu}$ is anti-self-dual. Therefore the positive chirality operator reduces to the scalar Klein-Gordon operator:
\begin{eqnarray}
m^2+D D^\dagger &=& m^2-D_\mu^2
\label{kg}
\end{eqnarray}
which implies that
\begin{eqnarray}
\Gamma_{\rm spinor}[A]=-2 \Gamma_{\rm scalar}[A]-\Delta\Gamma[A]
\label{spsc}
\end{eqnarray}
This familiar result enables the computation of the quark determinant in an instanton background via the associated scalar determinant, which has a partial-wave decomposition \cite{'tHooft:1976fv,Brown:1977bj,Carlitz:1978xu,Dunne:2004sx,Dunne:2005te}.
In this paper we study a special class of background gauge fields that are not self-dual, but for which the Dirac operator still admits a partial wave decomposition \cite{Adler:1972qq,Adler:1974nd,Bogomolny:1981qv,Bogomolny:1982ea,Fry:2003uy,Fry:2010cd}. Before coming to the radial decomposition, we first note that these gauge fields admit a simple chiral decomposition.
Specifically, we consider gauge fields of the form
\begin{eqnarray} \label{Apotential}
A_\mu (x) &=& \eta_{\mu \nu}^3 x_\nu g(r) \quad ,
\label{a}
\end{eqnarray}
where $\eta_{\mu \nu}^3$ are the 't Hooft symbols \cite{'tHooft:1976fv}, and $g(r)$ is a radial profile function, to be specified below.
This type of background field is symmetric under $O(2) \times O(3)$
transformations,
leading to a partial wave decomposition
\cite{Bogomolny:1981qv,Bogomolny:1982ea}. This decomposition applies for any radial profile function $g(r)$, so we have the freedom to study a wide class of background fields. In particular, we can investigate the role of zero modes, the presence or absence of which depends on the form of $g(r)$.
The associated field strength tensor is
\begin{equation}
F_{\mu \nu}(x) = - 2 \eta_{\mu \nu}^3 g(r) - \frac{g'(r)}{r} \left(\eta_{\mu \sigma}^3 x_\nu x_\sigma - \eta_{\nu \sigma}^3 x_\mu x_\sigma\right)
\label{f}
\end{equation}
Note that this field strength is not self-dual, due to the presence of the term proportional to $g'(r)$. Thus, the positive chirality operator $DD^\dagger$ does not simplify as in (\ref{kg}). However, for this field the negative chirality operator does take a particularly simple form:
\begin{eqnarray}
m^2+D^\dagger D=m^2-D_\mu^2+\left(4g(r)+r\,g'(r)\right)\sigma_3
\label{negative}
\end{eqnarray}
This is diagonal in spinor degrees of freedom, so if we work in the negative chirality sector, then we can immediately use our previous results for scalar Klein-Gordon operators \cite{Dunne:2004sx,Dunne:2005te,Dunne:2006ac,Dunne:2007mt,Hur:2008yg}, just by including an extra "potential" equal to $\pm \left(4g(r)+r\,g'(r)\right)$. Thus, we can write
\begin{eqnarray} \label{radial1}
m^2+D^\dagger D&=&-\left[ \partial_r^2 + \frac{4 l + 3}{r} \partial_r - r^2 g(r)^2 - 4 g(r) l_3 - m^2
\mp (4 g(r) + r g'(r) ) \right]
\nonumber\\
&\equiv&
m^2+\mathcal{H}_{(l, l_3, s)}
\end{eqnarray}
where the quantum number $l$ takes half-integer values: $l=0, \frac{1}{2}, 1, \frac{3}{2}, \dots$, while $l_3$ ranges from $-l$ to $l$, in integer steps. In this way, we compute $\Gamma^{(-)}[A]$, and then the full spinor effective action can then be obtained using (\ref{choice}) and (\ref{difference}).
\section{The partial-wave cutoff method}
In this section we briefly recall the partial-wave cutoff method developed previously for radially separable fields of the form (\ref{a}), adapted to the negative chirality sector of the spinor theory using (\ref{negative}). The basic idea of the partial-wave cutoff method involves separating the sum over the the quantum number $l$ into a low partial-wave contribution, each term of which is computed using the (numerical) Gelfand-Yaglom method, and a high partial-wave contribution, whose sum is computed analytically using a WKB expansion. The regularization and renormalization procedure tells us how to combine these two contributions to yield the finite and renormalized effective action \cite{Dunne:2004sx,Dunne:2005te,Dunne:2006ac,Dunne:2007mt,Hur:2008yg}.
\subsection{Low partial-wave contribution}
The low partial-wave contribution for our system is given by
\begin{eqnarray}
\Gamma^{(-)}_{{\rm{L}}} &=& -\sum_{s = \pm} \; \sum_{l =
0, \frac{1}{2} , 1, \ldots}^L \; \Omega(l) \sum_{l_3 = -l}^{l}
\ln \left( \frac{ {\rm{det}}(m^2+\mathcal{H}_{(l, l_3,
s)})}{{\rm{det}}(m^2+\mathcal{H}_{(l, l_3, s)}^{{\rm{free}}})} \right) \, ,
\end{eqnarray}
where $L$ is an arbitrary angular momentum cutoff.
The factor $
\Omega(l) = (2 l +1) $
is the degeneracy factor, and the
$s$ sum comes from adding the contributions of each spinor
component in \eqref{radial1}. In order to evaluate this quantity, we
use the Gel'fand-Yaglom method, which we now briefly review (for further details see \cite{Dunne:2004sx,Dunne:2005te,Dunne:2006ac}). Let
$\mathcal{M}_1$ and $\mathcal{M}_2$ denote two second-order radial
differential operators on the interval $r \: \in \, [\, 0,\infty)$
and let $\Phi_1(r)$ and $\Phi_2(r)$ be solutions to the following
initial value problem :
\begin{equation}
\mathcal{M}_i \Phi_i(r) = 0; \quad \Phi_i(r) \sim r^{2 l} \quad {\rm{as}} \quad r \to 0 \, .
\label{initial}
\end{equation}
Then the ratio of the determinants is given by
\begin{eqnarray}
\frac{{\rm{det}} \mathcal{M}_1}{{\rm{det}} \mathcal{M}_2} &=& \lim_{ R \to \infty} \left( \frac{\Phi_1(R)}{\Phi_2(R)} \right) \, .
\end{eqnarray}
In the negative chirality sector we can take, $\mathcal{M}_1 =m^2+\mathcal{H}_{(l, l_3, s)}$, and
$\mathcal{M}_2$ to be the corresponding free operator: i.e, the same operator with the background field set to zero: $g(r)\equiv 0$.
Thus, for a given radial profile function $g(r)$ in (\ref{a}), for each value of $(l, l_3, s)$ we need to solve
\begin{eqnarray}
\Phi_{\pm}''(r) + \frac{4 l + 3 }{r} \Phi_{\pm}'(r) - \left( m^2 +
4 l_3 g(r) + r^2 g(r)^2 \mp [4 g(r) + r g'(r)] \right)
\Phi_{\pm}(r) &=& 0 \, ,
\end{eqnarray}
with the initial value boundary condition in (\ref{initial}). The value of $\Phi$ at $r=\infty$ gives us the value of the determinant for that partial wave. In fact,
the corresponding free equation is analytically soluble.
It is numerically more convenient to define
\begin{eqnarray}
S_\pm^{(l, l_3 )}(r) &\equiv& \ln \left( \frac{\Phi_{l, l_3,\pm} (r)}{\Phi_{l, l_3,\pm}^{\rm free} (r)} \right) \, ,
\end{eqnarray}
and solve numerically the corresponding initial value problem for $S(r)$, as explained in \cite{Dunne:2005te,Dunne:2007mt}.
Then the contribution of the low-angular-momentum partial-waves to the effective action is
\begin{eqnarray}
\Gamma^{(-)}_{{\rm{L}}} &=& - \sum_{l = 0, \frac{1}{2}, 1 ,
\ldots}^L \; \Omega(l) \sum_{l_3 = -l}^{l} [ S_{+}^{(l,
l_3)}(\infty) + S_{-}^{(l, l_3)}(\infty) ] \, .
\end{eqnarray}
While each term in the sum over $l$ is finite and simple to compute, the sum over $l$ is divergent as $L \to \infty$. In fact, only after
adding the renormalized contribution of the high partial-wave modes do we obtain a finite and renormalized result for the effective action.
\subsection{High partial-wave contribution}
The high partial-wave contribution for our system is given by
\begin{eqnarray} \label{gammalow}
\Gamma^{(-)}_{{\rm{H}}} &=& -\sum_{s = \pm \frac{1}{2}} \; \sum_{l = L
+ \frac{1}{2}}^\infty \; \Omega(l) \sum_{l_3 = -l}^{l} \ln
\left( \frac{ {\rm{det}}(m^2+\mathcal{H}_{(l, l_3, s)})}{{\rm{det}}(m^2+\mathcal{H}_{(l, l_3, s)}^{{\rm{free}}} )}
\right) \, .
\end{eqnarray}
Again, since by (\ref{negative}) we only need the negative chirality sector, and the negative chirality sector diagonalizes fully, we can apply our previous scalar analysis to each of the diagonal components, adding the appropriate term $\pm (4g+rg')$ to the Klein-Gordon operator. This modifies the detailed expressions as follows. In the large $L$ limit, we have
\begin{align}
\Gamma^{(-)}_{{\rm{H}}} &= \int_0^\infty dr \left( \frac{8 \, g(r) r^3}{3 \sqrt{\tilde{r}^2 + 4}} \right) L^2
+ \int_0^\infty dr \left( \frac{2 r^3 ( 3 \tilde{r}^3 + 8) g(r)^2}{( \tilde{r}^2 + 4)^{3/2}} \right) L \\
&+ \int_0^\infty dr \Bigg\{ \frac{r^3}{45 ( \tilde{r}^2 +
4)^{7/2}} \Bigg[ -6 r^4 (5 \tilde{r}^4 + 28 \tilde{r}^2 + 32 )g(r)^4
\\ &+ 15 (33 \tilde{r}^6 + 335 \tilde{r}^4 + 1192 \tilde{r}^2 + 1600) g(r)^2 \\
&+ 10 r (15 \tilde{r}^6 + 184 \tilde{r}^4 + 776 \tilde{r}^2 + 1120) g(r)g'(r) \\
&+ 5 r^2 (3 \tilde{r}^6 + 38 \tilde{r}^4+ 160 \tilde{r}^2 + 224) g'(r)^2 + 20 r^2 (4 + \tilde{r}^2)^2
g(r)g''(r)]
\Bigg] \\
&+ \frac{r^3 (20 g(r)^2 + 10 g(r)g'(r)r + g'(r)^2 r^2 ) }{12} \Bigg[
\gamma + 2 \ln L - 2 \ln \Bigg( \frac{r}{2 + \sqrt{\tilde{r}^2 +
4}} \Bigg)
\Bigg]
\Bigg\} \\
&- \frac{r^3 (20 g(r)^2 + 10 g(r)g'(r) r + g'(r)^2 r^2 )}{12 \,
\epsilon} + O\Bigg( \frac{1}{L} \Bigg) \, ,
\end{align}
where $\tilde{r} \equiv \frac{r m}{L}$. We now identify the
counterterm as
\begin{equation}
\delta \Gamma^{(-)} = \frac{1}{12} \Bigg( \frac{1}{\epsilon} - \gamma - 2
\ln \mu \Bigg) \int_0^\infty dr \; r^3 (20 g(r)^2 + 10 g(r)g'(r) r +
g'(r)^2 r^2 ) \, ,
\label{ct}
\end{equation}
Where $\mu$ is the renormalization scale.
Note that
\begin{eqnarray}\label{FF}
\frac{1}{2} \int_0^\infty dr \; r^3 (F_{\mu \nu} F_{\mu \nu}) &=&
\int_0^\infty dr \; r^3 (8 g(r)^2 + 4 g(r)g'(r) r +
g'(r)^2 r^2 ) \, , \\
\frac{1}{2} \int_0^\infty dr \; r^3 (F_{\mu \nu} \tilde{F}_{\mu \nu}) &=&
\int_0^\infty dr \; r^3 (8 g(r)^2 + 4 g(r)g'(r) r ) \, ,
\label{FFdual}
\end{eqnarray}
and thus, the counterterm corresponds to the following combination:
\begin{equation}
\delta \Gamma^{(-)} = \frac{1}{24} \Bigg( \frac{1}{\epsilon} - \gamma - 2
\ln \mu \Bigg) \int_0^\infty dr \; r^3 \Bigg( F_{\mu \nu} F_{\mu \nu} + \frac{3}{2} F_{\mu \nu} \tilde{F}_{\mu \nu} \Bigg) \, .
\end{equation}
The appearance of the $F_{\mu \nu} \tilde{F}_{\mu \nu}$ term
is a special feature of the spinor calculation that does not occur in the scalar case.
Having identified the counter-term, we can now write an explicit expression for the large $L$ behavior of the high partial-wave contribution to the renormalized effective action:
\begin{eqnarray}
\Gamma_{{\rm{H}}}^{(-)} = \int_0^\infty dr \left( Q_{{\rm {log}}}(r)\ln L + \sum_{n = 0}^2 Q_n ( r ) L^n + \sum_{n = 1}^N Q_{-n}( r ) \frac{1 }{ L^n } \right) + O(\frac{1}{L^{n+1}}) \, ,
\label{gammaren}
\end{eqnarray}
with the following expansion coefficients:
\begin{align}
Q_2(r) &= \frac{8 \, g(r) r^3}{3 \sqrt{\tilde{r}^2 + 4}} \nonumber \\
Q_1(r) &= \frac{2 r^3 (3 \tilde{r}^3 + 8) g(r)^2}{(
\tilde{r}^2 + 4)^{3/2}} \nonumber \\
Q_{{\rm {log}}}(r) &= -\frac{1}{6}\, r^3 (20 g(r)^2 + 10 g(r)g'(r) r + g'(r)^2
r^2 ) \nonumber \\
Q_0(r) &= \frac{r^3}{45 ( \tilde{r}^2 +
4)^{7/2}} \Bigg[ -6 r^4 (5 \tilde{r}^4 + 28 \tilde{r}^2 + 32 )g(r)^4
\nonumber \\ &+ 15 (33 \tilde{r}^6 + 335 \tilde{r}^4 + 1192 \tilde{r}^2 + 1600) g(r)^2 \nonumber \\
&+ 10 r (15 \tilde{r}^6 + 184 \tilde{r}^4 + 776 \tilde{r}^2 + 1120) g(r)g'(r) \nonumber \\
&+ 5 r^2 (3 \tilde{r}^6 + 38 \tilde{r}^4+ 160 \tilde{r}^2 + 224) g'(r)^2 + 20 r^2 (4 + \tilde{r}^2)^2
g(r)g''(r) \Bigg]
\nonumber \\
&- Q_{{\rm {log}}}(r) \ln \Bigg( \frac{ \mu r}{2 + \sqrt{\tilde{r}^2 + 4}}
\Bigg)
\nonumber \\
Q_{(-1)}(r) &= -\frac{r^3}{4 ( \tilde{r}^2 + 4)^{9/2}} \Bigg[ 6
r^4( \tilde{r}^6 + 4 \tilde{r}^4 )g(r)^4 \nonumber\\
&+2 r^2( \tilde{r}^6 + 16 \tilde{r}^4 + 80\tilde{r}^2 +
128)g'(r)^2 \nonumber \\
&+ (-4 \tilde{r}^8 + 89 \tilde{r}^6 + 1104 \tilde{r}^4 + + 3456 \tilde{r}^2 + 5120 ) g(r)^2 \nonumber \\
&+16 r( 2 \tilde{r}^6 + 21 \tilde{r}^4 + 92
\tilde{r}^2 + 160)g(r) g'(r) \nonumber \\
&-4 r^2(\tilde{r}^6 + 8 \tilde{r}^4 + 16
\tilde{r}^2 )g(r) g''(r)
\Bigg] \, . \label{spinorQs}
\end{align}
Note that $\Gamma_{{\rm{H}}}^{(-), \rm{ren}}$ involves $L^2$, $L$ and
$\ln L$ terms that diverge as $L \to \infty$, but these divergences
exactly cancel those of the $\Gamma_{{\rm{L}}}$ contribution (coming from the numerical
Gel'fand-Yaglom computation for the low partial-wave modes),
yielding a finite renormalized result for the effective action. This method is essentially
exact (up to numerical precision) and accurately provides the value of the effective action
for any value of the mass:
\begin{equation}
\Gamma^{(-)}_{\rm{ren}} (m) = \Gamma_{{\rm{H}}}^{(-), \rm{ren}} (m)+ \Gamma^{(-)}_{{\rm{L}}} (m) \quad .
\end{equation}
The effective action calculated as above is finite for any non-zero value of the mass, however, from eq.~(\ref{spinorQs}), we note that $Q_0 (r)$ contains a term proportional
to $\ln \mu$ and therefore when we use "on-shell" renormalization ($\mu = m$), we have
\begin{equation}
\Gamma^{\rm{ren}} (m) \sim \Bigg( - \int_0^\infty dr \, Q_{{\rm {log}}}(r) \Bigg) \ln m \quad \quad ; \quad \quad m \to 0 \, .
\end{equation}
In order to analyse the mass dependence of the effective action, we introduce
a modified effective action defined as
\begin{equation}
\tilde{\Gamma}^{\rm{ren}} (m) \equiv \Gamma^{\rm{ren}} (m) + \Bigg( \int_0^\infty dr \, Q_{{\rm {log}}}(r) \Bigg) \ln \mu \,.
\end{equation}
which is independent of the renormaliztion scale $\mu$, and which is
finite at $m=0$.
\section{Results for specific background profiles } \label{results}
Within the class of background gauge fields (\ref{Apotential}), we still have the freedom to specify the radial profile function $g(r)$. Here we note that the large $r$ behavior of $g(r)$
determines the presence or absence of zero-modes. To be specific,
the integral of $F_{\mu \nu} \tilde{F}_{\mu \nu}$ counts the
number of zero-modes. From (\ref{FFdual}) we have:
\begin{eqnarray}
\frac{1}{2} \int_0^\infty dr \; r^3 (F_{\mu \nu} \tilde{F}_{\mu \nu}) &=&
\int_0^\infty dr \; r^3 (8 g(r)^2 + 4 g(r)g'(r) r ) \, , \nonumber \\
&=& 2 (g(r) r^2)^2 \Big\vert_0^{\infty} \, .
\end{eqnarray}
Therefore, as long as $g(r)$ falls faster than $1/r^2$ we don't have
zero-modes. In this paper, we study the background fields using two different
profile functions to set the background field. The first one is
\begin{equation} \label{g1}
g_1 (r) \equiv B (1 - {\rm {Tanh}}( \beta (\sqrt{B} r - \xi ) )) \, ,
\end{equation}
where $B$, $\beta$ and $\xi$ are parameters that control the amplitude, range and steepness of the potential. We may call $g_1(r)$ the "step potential". Since $g_1(r)$
falls off exponentially fast as $r \to \infty$, this case
is free of zero-modes.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.6\textwidth]{fig1a.pdf}
\includegraphics[width=0.6\textwidth]{fig1b.pdf}
\end{center}
\caption{ The upper figure shows the radial profile function: $g_1(r) = B (1 - {\rm {Tanh}}( \beta (\sqrt{B} r - \xi ) )
)$. We have fixed $B=1$ and $\beta=1$ in all the curves. The parameter $\xi$ which controls the \emph{range}
of the potential is varied. The lowest curve corresponds to $\xi=1$ (solid/red),
followed by curves corresponding to $\xi=3/2$ (dashes/blue), $\xi=2$ (dots/black), $\xi=3$ (dot-dashes/green).
The lower figure shows a second radial profile function: $g_2 (r) = \nu e^{- \alpha r^2}/(\rho^2 + r^2)$. We have fixed $\nu=1$ and $\rho=1$ in all the curves, the lowest curve corresponds to $\alpha = 1$ (solid/red),
followed by curves corresponding to $\alpha = 1/20$ (dashes/blue), and $\alpha = 1/400 $ (short-dashes/black). }
\label{fig1}
\end{figure}
The second profile function we use is
\begin{equation} \label{g2}
g_2 (r) \equiv \nu\, e^{-\alpha r^2}/(\rho^2 + r^2) \, ,
\end{equation}
where $\nu$,
$\rho$ and $\alpha$ are parameters that control the amplitude, range and steepness of the potential. This potential has the following properties:
\begin{eqnarray}
\alpha > 0 &\Longrightarrow& \quad \int d^4 x F_{\mu \nu} F_{\mu \nu} < \infty
\quad , \quad \int d^4 x F_{\mu \nu} \tilde{F}_{\mu \nu} = 0 \nonumber \\
\alpha = 0 &\Longrightarrow& \quad \int d^4 x F_{\mu \nu} F_{\mu \nu} \to \infty
\quad , \quad 0 < \Big| \int d^4 x F_{\mu \nu} \tilde{F}_{\mu \nu} \Big| < \infty \, .
\end{eqnarray}
Note that $g_2(r)$ goes like $1/r^2$
when we set $\alpha = 0$, so that in this case there are zero modes. However, whenever the profile function $g(r)$ decays as $1/r^2$ we have
\begin{eqnarray}
\frac{1}{2} \int_0^\infty dr \; r^3 F_{\mu \nu} F_{\mu \nu} &=&
\int_0^\infty dr \; r^3 (8 g(r)^2 + 4 g(r)g'(r) r +
g'(r)^2 r^2 ) \, ,\nonumber \\
&=& \int_0^\infty dr \; r^3 F_{\mu \nu} \tilde{F}_{\mu \nu} + \int_0^\infty dr \; r^3 ( g'(r)^2 r^2 )
\end{eqnarray}
which diverges logartihmically.
In Figure \ref{fig1} we show some plots of our chosen profile functions
$g_1(r)$ and $g_2(r)$ for different values of their parameters.
Our calculational method allows one to compute the effective action for arbitrary values of the fermion mass $m$.
This provides us with a unique opportunity to study the validity of various approximate methods that yield estimates in the limits of large or small mass. In particular, we are able to probe exactly the $m\to 0$ limit, which is of great interest in spinor theories but which is
notoriously inaccessible by approximate means.
\section{Approximation methods} \label{approximations}
\subsection{The large-mass expansion}
The large-mass expansion is the most general approximation method in the
sense that that it may be applied to calculate the effective action
for any well-behaved background. Its main limitation is that it only
applies for large values of the mass and, in fact, it diverges as
$m \to 0$. Thus, it is not directly useful for probing issues related to massless quarks.
To outline this method, consider for instance, the spinor
case
\begin{equation}
\Gamma =
-\frac{1}{2}\ln {\rm {det}} ( - {\displaystyle{\not} D}^2 + m^2) \nonumber \\
= \frac{1}{2}\int_0^{\infty} \frac{ds}{s} e^{- m^2 s }\, {\rm {Tr}}\, e^{- s (-{\displaystyle{\not}
D}^2)} \, .
\end{equation}
In order to analyse the large-mass limit, we may take the small $s$
limit in the proper-time integral and expand the trace factor in
powers of $s$, yielding the heat kernel or Seeley-De Witt expansion:
\begin{equation}
s \to 0 \quad : \quad {\rm {Tr}}\, e^{- s (-{\displaystyle{\not}
D}^2)} \sim \frac{1}{(4 \pi s )^{d/2}} \sum_{n = 0}^{\infty} s^n a_n [
F_{\mu \nu} ] \, .
\end{equation}
The Seeley-DeWitt coefficients $a_n$ are given by Lorentz traces of
powers of $F_{\mu \nu }$and its derivatives.
For the $A_{\mu} (x)$ backgrounds given by (\ref{g1}) and (\ref{g2}) the Seeley-DeWitt coefficients are proportional to those found in the scalar case:
\begin{eqnarray}
a_1 &=& 0 \, , \nonumber \\
a_2 &=& \frac{2}{3} \int d^4x\, F_{\mu \nu} F_{\mu \nu}
\, , \nonumber \\
a_3 &=& -\frac{ 3}{45} \int d^4x \, (D_\mu F_{\nu
\lambda})(D_\mu F_{\nu \lambda}) \, .
\end{eqnarray}
Since we are
considering potentials of the form (\ref{Apotential}), with field strength (\ref{f}), we find an expansion of the renormalized effective action as:
\begin{equation}
\tilde{\Gamma}^{{\rm ren}} = \tilde{\Gamma}_{{\rm LM}}^{(0)} \ln m + \tilde{\Gamma}_{{\rm LM}}^{(2)} \frac{1}{m^2} + \tilde{\Gamma}_{{\rm LM}}^{(4)} \frac{1}{m^4} + \cdots
\end{equation}
The first two coefficients are:
\begin{eqnarray}
\tilde{\Gamma}_{{\rm LM}}^{(0)} &=& \frac{1}{6} \int_0^\infty dr \; r^3 (8 g(r)^2 + 4 g(r)g'(r) r +
g'(r)^2 r^2 ) \, , \nonumber \\
\tilde{\Gamma}_{{\rm LM}}^{(2)} &=& \frac{1}{180} \int_0^\infty dr \big[ 24 r^2 g(r) \big( 15 g'(r) + r \big( 9 g''(r) + r g^{(3)}(r) \big) \big) \nonumber \\
&\, & + r^3 \big( 221 g'(r)^2 + 9 r^2 g''(r)^2 + 2 r g'(r) \big( 71 g''(r) + 6 r g^{(3)}(r) \big) \big) ]
\end{eqnarray}
Figure \ref{fig2} presents a comparison between the effective action, as calculated using large-mass expansion and its
exact value obtained using our partial-wave cutoff method. As expected, both methods agree well for large masses.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.75\textwidth]{fig2.pdf}
\end{center}
\caption{ The graph shows the one-loop effective action, $\tilde{\Gamma} (m)$, as a function of fermion mass $m$, with the radial profile function: $g_1(r) = B (1 - {\rm
{Tanh}}( \beta (\sqrt{B} r - \xi ) ))$. We have set $B=1$, $\beta=1$ and $\xi=1$. The
dots correspond to our exact numerical results, based on the Gelfand-Yaglom theorem, $\tilde{\Gamma}_{{\rm GYT}}(m)$, and the solid line shows
the outcome of the larg-mass expansion $\tilde{\Gamma}_{{\rm LM}}(m)$. }
\label{fig2}
\end{figure}
\subsection{The derivative expansion}
Another widely used approximation in gauge theories is the derivative expansion. This method is based on the fundamental result that for background gauge fields with constant field strength $F_{\mu\nu}$ it is possible to compute the renormalized effective action in a simple analytic form \cite{Heisenberg:1935qt,Schwinger:1951nm,Dunne:2004nc}. One can then expand around this soluble case, leading to an expansion of the form \cite{Cangemi:1994by,Gusynin:1995bc,Salcedo:2000hp}
\begin{equation}
S_{{\rm eff}} \approx S_0 [F] + S_2 [F, (\partial F)^2] + S_4 [F, (\partial F)^2, (\partial F)^4 ] + \cdots
\label{der}
\end{equation}
The leading term in this expansion is the well-known one-loop effective action for constant backgrounds, first computed by Heisenberg and Euler \cite{Heisenberg:1935qt}.
In euclidean QED, the corresponding one-loop effective {\it Lagrangians} for spinor and scalar QED are
\begin{eqnarray}
\mathcal L_{{\rm spinor}}( a, b)
&=& -\frac{1}{8 \pi^2} \int_0^{\infty} \, \frac{ds}{s^3}\, e^{-m^2 s }
\Big\{ a b s^2\, {\rm coth}(a s) {\rm coth}(b s)
+ 1 - \frac{ s^2}{3}\,( a^2 + b^2 ) \Big\} \, , \label{euclidean-EHsp}
\end{eqnarray}
and
\begin{eqnarray}
\mathcal L_{{\rm scalar}}( a, b)
&=& \frac{1}{16 \pi^2} \int_0^{\infty} \, \frac{ds}{s^3}\, e^{-m^2 s }
\Bigg\{ \, \frac{ a b\, s^2}{{\rm sinh}( a s) {\rm sinh}( b s)}
- 1 + \frac{ s^2}{6}\,( a^2 + b^2 ) \Bigg\} \label{euclidean-EHsc} \,,
\end{eqnarray}
where $\pm i a$ and $\pm b$ are the eigenvalues of the $4 \times 4$ constant matrix $F_{\mu \nu}$, and are related to the field invariants in the following way:
\begin{equation}
a^2 + b^2 = \frac{1}{2} F_{\mu \nu} F^{\mu \nu} \quad , \quad
a b = \frac{1}{2} F_{\mu \nu} \tilde{F}^{\mu \nu} \, .
\end{equation}
The leading term in the derivative expansion is obtained by simply replacing the
the constants $a$ and $b$ by the corresponding space dependent quantities inside these integral expressions, and then obtaining the effective action from the effective Lagrangian by integrating over spacetime. For the class of backgrounds that we are considering, the corresponding substitution is
\begin{eqnarray}
a &\longrightarrow& a(r) = 2 g(r) \, , \nonumber \\
b &\longrightarrow& b(r) = 2 g(r) + r g'(r) \, ,
\end{eqnarray}
such that,
\begin{equation}
\Gamma_{\rm DE} = 2 \pi^2 \int_0^\infty r^3\,dr \, \mathcal{L} (a(r), b(r)) \, .
\end{equation}
At first sight, one would expect the derivative expansion to be a good approximation at large mass, similar to the large mass expansion, since the expansion over derivatives in (\ref{der}) is balanced by inverse powers of $m$. However, the situation is more subtle than that naive expectation, as the derivative expansion expression at a given order is a resummation of all powers of the field strength with a fixed number of derivatives. Thus, one might expect that the derivative expansion is better than the large mass expansion for smaller values of the mass. Indeed, in previous work on the partial-wave cutoff method in scalar theories it was found that even
the leading term of the derivative expansion provides a surprisingly accurate approximation to the mass dependence of effective action~\cite{Dunne:2007mt}, even for very small mass. In the next section we study this question for spinor theories and find an interesting difference.
\section{Approximation methods versus exact calculation}
In this section we use exact results obtained from our partial-wave cutoff method to probe the validity of the large-mass and deriative expansion approximations, for spinor QED.
In order to emphasize
the similarities and differences between the scalar and spinor cases, we also show some results for the
scalar action as well.
\subsection{The scalar case}
To exemplify how the partial-wave cutoff method works in the scalar theory, we use the background field given by $g_1(r)$, choosing different values of the range parameter $\xi$. We present
a comparison between the the exact effective
action as calculated with the partial-wave cutoff method, the large-mass expansion and the derivative
expansion.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.6\textwidth]{fig3a.pdf}
\includegraphics[width=0.6\textwidth]{fig3b.pdf}
\includegraphics[width=0.6\textwidth]{fig3c.pdf}
\end{center}
\caption{ The graphs show, for the case of scalar QED, our exact effective action, $\tilde{\Gamma}_{{\rm GYT}}(m)$ (big-dots/red), the derivative expansion expression $\tilde{\Gamma}_{{\rm DE}}(m)$ (small-dots/blue), and the large mass expansion expression, $\tilde{\Gamma}_{{\rm LM}}(m)$ (solid-line).
We use the profile function $g_1(r)= B (1 - {\rm
{Tanh}}( \beta (\sqrt{B} r - \xi ) ))$ with $B = 1$, $\beta = 1$. We show results for three values of the \emph{range} parameter: $\xi = 1$ (upper graph), $\xi = 3/2$ (middle graph) and $\xi = 2$ (lower graph). Notice that the derivative expansion is a reasonable approximation even at zero mass. }
\label{fig3}
\end{figure}
From the plots in Figure \ref{fig3}, we make the following observations:
\begin{itemize}
\item The large-mass expansion result agrees very well with the exact result, already for $m \sim 2$.
\item The leading order derivative expansion is good at large $m$, and also provides a very good approximation
to the effective action for large values of the parameter $\xi$.
\item In particular, as long as the steepness parameter $\xi$ is large enough, the derivative expansion provides
accurate results in both the large-mass and small-mass regimes.
\end{itemize}
\subsection{The spinor case}
In this section we compare and analyse the different approximation methods against the GYT method. The first background configuration we investigate is $g_1(r)$, the plots
shown in fig.~(\ref{fig4}) correspond to the GYT method, the large-mass expansion and the derivative expansion.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\textwidth]{fig4.pdf}
\end{center}
\caption{For spinor QED, this plot shows the exact effective action, $\tilde{\Gamma}_{{\rm GYT}}(m)$ (big-dots/red), the derivative expansion expression, $\tilde{\Gamma}_{{\rm DE}}(m)$ (small-dots/blue), and the large mass expansion expression, $\tilde{\Gamma}_{{\rm LM}}(m)$ (solid-line). We use the same profile function as in the upper plot of Figure \ref{fig3} [which is for scalar QED]: $g_1(r)= B (1 - {\rm
{Tanh}}( \beta (\sqrt{B} r - \xi ) ))$ with $B = 1$, $\beta = 1$ and $\xi = 1$. Note the very different behavior at small mass compared to the scalar QED case.}
\label{fig4}
\end{figure}
From the plot in Figure \ref{fig4}, we make the following observations:
\begin{itemize}
\item The partial-wave cutoff method produces a finite value for the effective action in the small-mass regime.
\item As in scalar QED, for spinor QED the large-mass expansion result agrees very well with the exact result, already for $m \sim 2$.
\item As in scalar QED, for spinor QED the leading order derivative expansion is good at large $m$. However, the leading order derivative expansion result diverges in the small-mass regime.
\end{itemize}
To understand why the leading-order derivative expansion fails in the small-mass limit, we analyse equation
(\ref{euclidean-EHsp}) in the small-mass regime. The small-mass regime is obtained by taking $s \to \infty$, which gives:
\begin{eqnarray}
\mathcal{L}_{{\rm spinor}}( a, b)
&\sim& -\frac{1}{8 \pi^2} \Big[ a b - \frac{ 1}{3}\,( a^2 + b^2 ) \Big] \int_0^{\infty} \, \frac{ds}{s}\, e^{-m^2 s }
\nonumber \\
&\sim& \frac{1}{4 \pi^2} \Big[ a b - \frac{ 1}{3}\,( a^2 + b^2 ) \Big] \ln m \quad .
\label{spinorde}
\end{eqnarray}
We recognize one $\ln m$ divergence proportional to $F_{\mu\nu}F_{\mu\nu}$, and another proportional to $F_{\mu\nu}\tilde{F}_{\mu\nu}$. Contrast this with scalar QED where
\begin{eqnarray}
\mathcal{L}_{{\rm scalar}}( a, b)
&\sim& \frac{1}{16 \pi^2} \Big[ 0 + \frac{ 1}{6}\,( a^2 + b^2 ) \Big] \int_0^{\infty} \, \frac{ds}{s}\, e^{-m^2 s }
\nonumber \, , \\
&\sim& - \frac{1}{8 \pi^2} \Big[ \frac{ 1}{6}\,( a^2 + b^2 ) \Big] \ln m \quad .
\end{eqnarray}
which has a $\ln m$ divergence proportional to $F_{\mu\nu}F_{\mu\nu}$, but none proportional to $F_{\mu\nu}\tilde{F}_{\mu\nu}$. This is simply a reflection of the fact that zero modes can occur in spinor QED but not in scalar QED, with the $F_{\mu\nu}\tilde{F}_{\mu\nu}$ term determining the number of zero modes.
But note that $\mathcal{L}_{{\rm spinor}}( a, b) = \mathcal{L}_{{\rm spinor}}( -a, -b)$. Therefore, (\ref{spinorde}) really should read
\begin{eqnarray}
\mathcal{L}_{{\rm spinor}}( a, b)
\sim \frac{1}{4 \pi^2} \Big[ |a b| - \frac{ 1}{3}\,( a^2 + b^2 ) \Big] \ln m \quad .
\label{spinorde2}
\end{eqnarray}
Therefore, our naive application of the leading-order derivative expansion includes a term in the effective action that behaves in the small mass limit as
\begin{eqnarray}
\left(\int d^4 x\, |a(r)b(r)|\right)\ln m
\label{1}
\end{eqnarray}
instead of the correct form
\begin{eqnarray}
\left(\int d^4 x\, a(r)b(r)\right)\ln m
\label{2}
\end{eqnarray}
It is this latter form that counts the number of zero modes, and appears in the counter-term. To show the effect of this, consider the radial profile function $g_1(r)$ for which $\int d^4 x\, F_{\mu\nu}\tilde{F}_{\mu\nu}=0$, indicating the absence of zero modes. For this profile function, this integral vanishes because the integrand changes sign. But the derivative expansion expression does not allow for such changes of sign, and so we effectively compute $\int d^4 x\, \left | F_{\mu\nu}\tilde{F}_{\mu\nu}\right |$, which is non-zero.
This mis-match is demonstrated clearly in Figures \ref{fig5}, which show that the divergent small-mass behavior of the derivative expansion corresponds exactly to
\begin{equation}
\tilde{\Gamma}_{{\rm DE}}^{{\rm spinor}}(m) \sim f(m) = \Big( \frac{1}{4 \pi^2} \int d^4 x \; |a(r)b(r)| \Big) \, \ln m \quad ; \quad m \to 0 \, .
\end{equation}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.6\textwidth]{fig5a.pdf}
\includegraphics[width=0.6\textwidth]{fig5b.pdf}
\includegraphics[width=0.6\textwidth]{fig5c.pdf}
\end{center}
\caption{For spinor QED, these graphs show the exact effective action, $\tilde{\Gamma}_{{\rm GYT}}(m)$ (big-dots/red), the derivative expansion expression, $\tilde{\Gamma}_{{\rm DE}}(m)$ (small-dots/blue), and the large-mass expansion expression $\tilde{\Gamma}_{{\rm LM}}(m)$ (solid-line). The dashed line represents the residual logarithm $f(m) = \frac{1}{4 \pi^2} (\int d^4 x |a(r)b(r)|) \ln m$. We use the profile function $g_1(r)= B (1 - {\rm
{Tanh}}( \beta (\sqrt{B} r - \xi ) ))$ with $B = 1$, $\beta = 1$. We show results for three values of the \emph{range} parameter: $\xi = 1$ (upper graph), $\xi = 3/2$ (middle graph) and $\xi = 2$ (lower graph). Note that the divergent behavior at small mass is well fitted by the residual logarithm, as described in the text.}
\label{fig5}
\end{figure}
The number of zero-modes is given by $\frac{1}{4 \pi^2} \int d^4 x a(r) b(r) $ and it is always equal to zero for the
backgrounds of the class $g_1(r)$, however, $\frac{1}{4 \pi^2} \int d^4 x |a(r)b(r)|$ does not vanish and this adds an
incorrect residual logarithmic dependence to the derivative expansion, as we have shown.
No such mis-match occurs for scalar QED because the small $m$ behavior of the derivative expansion expression only involves the combination $F_{\mu\nu}F_{\mu\nu}\propto [a^2(r)+b^2(r)]$, which is always positive. Thus there is no $\ln m$ divergence in the small $m$ behavior of the derivative expansion plots shown in Figure \ref{fig3}
for scalar QED.
The second background we examined is the one given by the profile function $g_2(r)$.
Setting different values of the parameter $\alpha$, we can control the rate of decay
of the potential. In figure~\ref{fig6}, we present a comparison between the different calculation
methods for this kind of background.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.6\textwidth]{fig6a.pdf}
\includegraphics[width=0.6\textwidth]{fig6b.pdf}
\includegraphics[width=0.6\textwidth]{fig6c.pdf}
\end{center}
\caption{ For spinor QED, these graphs show the exact effective action, $\tilde{\Gamma}_{{\rm GYT}}(m)$ (big-dots/red), the derivative expansion expression, $\tilde{\Gamma}_{{\rm DE}}(m)$ (small-dots/blue), and the large-mass expansion expression $\tilde{\Gamma}_{{\rm LM}}(m)$ (solid-line). The dashed line corresponds to $\frac{1}{4 \pi^2} (\int d^4 x | a(r) b(r)|) \ln m$.
We use the profile function $g_2 (r) = \nu e^{-\alpha r^2}/(\rho^2 + r^2)$ with $\nu = 1$ and $\rho = 1$. We show three values
of the decay rate parameter: $\alpha = 1$ (upper graph), $\alpha = 1/20$ (middle graph) and $\alpha = 1/400$ (lower graph). Note that the divergent behavior at small mass is well fitted by the residual logarithm, as described in the text.}
\label{fig6}
\end{figure}
We corroborate once more how the derivative expansion fails at the zero-mass limit. Also, the derivative expansion shows better accuracy in a wider mass-range as the
for those fields with slower variation (small $\alpha$), as expected.
It also evident that the residual logarithm is not the dominant term in the derivative expansion when we move from the small-mass regime in to the large-mass regime.
\section{Conclusion}
We have extended the partial-wave
cutoff method to spinor theories with nontrivial [and non-self-dual] radially symmetric gauge backgrounds. Different background fields were tested, resulting always in accurate values of the effective action for both
the large-mass and small-mass regimes. We provided an example of how our method allows one to systematically
investigate how the effective action responds to different characteristics of the background fields, such as range, amplitude or rate of variation.
We have also analyzed how certain approximation methods compare with these exact results, in
the different mass regimes, also comparing the scalar case with the the spinor case. Specifically, we have tested the large-mass expansion and the derivative expansion. We have shown that the large mass expansion works extremely well, as in the scalar case. However, the derivative expansion behaves in a different manner in the small-mass regime for the spinor theory. We have explained this fact, both qualitatively and quantitatively, as resulting from the changing sign of the local quantity $F_{\mu\nu}\tilde{F}_{\mu\nu}$, and which in the derivative expansion approximation is assumed to be constant and therefore of fixed sign. We have also shown that this is directly related to the appearance of fermion zero modes.
\bigskip
GD and AH were supported by the US DOE grant DE-FG02-92ER40716, and HM was supported by the National Research
Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (No 2010-0011223).
|
\section{Introduction}
Understanding hadron structure from first principles is a fundamental
goal of lattice QCD. The nucleon plays a special role as a benchmark
for lattice calculations due to the extensive experimental effort to
measure its properties. Successfully reproducing the measured values
of basic observables, like the axial charge measured in neutron beta
decay or the charge radius measured in elastic electron scattering,
would provide a strong validation of the lattice technique. This
would not only give confidence in the calculations of the many other
properties of the nucleon but it would also bolster the lattice effort
to calculate hadron structure more generally. Furthermore, there is a
burgeoning program to calculate nuclear properties using lattice QCD.
It is clearly essential to have a well-controlled calculation of the
single nucleon state in order to trust future computations of
multi-nucleon systems. Thus in many ways the nucleon is the keystone
for a much broader lattice QCD program to understand the properties of
hadrons as predicted from the underlying theory of QCD.
I have chosen to illustrate the status of the lattice effort to
understand nucleon structure by focusing on average-$x$. This quantity
has persistently come out too high from lattice calculations. A
variety of explanations have been offered over the years, and I'll
mention a few, but recent calculations have dramatically confirmed
this trend. The apparent disagreement with experiment is a real
puzzle and its resolution will likely require a concentrated effort to
carefully examine all sources of error in the lattice calculation.
The advantage of using lattice QCD, as opposed to any other technique,
to calculate average-$x$ is that the list of possible errors is finite
and each source of error can be systematically removed. This is both
a challenge and an opportunity for lattice QCD.
\section{Average-$x$}
By average-$x$, I mean specifically the difference of the up and down
contributions. Written as a moment of the nucleon parton
distributions, average-$x$ is
\begin{equation}
\label{moment}
\langle x \rangle^{u-d}_\mu = \int_{0}^{1}\!\! dx\,\, x\, (u(x,\mu) - d(x,\mu)) + \int_{0}^{1}\!\! dx\,\, x\, (\overline{u}(x,\mu) - \overline{d}(x,\mu))\,.
\end{equation}
The unpolarized quark and anti-quark distribution functions $q(x,\mu)$
for $q=u$, $d$, $\overline{u}$ and $\overline{d}$ are extracted from
the results of many experiments, particularly deeply inelastic
electron-nucleon scattering. It is important to remember that the
various $q(x,\mu)$ are indirectly determined by performing global fits
to the measured cross-sections and that the values of $x$ are limited
by the kinematics of each experiment. For this reason, it would be
preferable to calculate $q(x,\mu)$ directly as a function of $x$, but
this is not possible with the lattice QCD methods that we currently
have. This is because lattice computations are performed in Euclidean
space whereas the distributions $q(x,\mu)$ are related to the square of the light-cone
wave function, which are not easily accessible outside of Minkowski
space-time. However, moments of the quark distributions can be
related to matrix elements of local operators, and these are
calculable in Euclidean space. Thus lattice computations determine
$\langle x \rangle^{u-d}_\mu$ from
\begin{equation}
\label{me}
\langle p, s | \left. \left( \overline{u} \gamma_{\{ \mu } iD_{\nu \}}u - \overline{d} \gamma_{\{ \mu } iD_{\nu \}}d \right)\right|_\mu | p, s \rangle = 2 \langle x \rangle^{u-d}_\mu p_{\{\mu} p_{\nu\}}\,.
\end{equation}
As we begin to contemplate what calculations are required for a
definitive determination of average-$x$, it is important to keep in mind
that there is a significant difference between what is computed and
how that is measured. Currently, the burden is on the lattice
community to nail down the calculations of the nucleon matrix
elements, but it is not inconceivable that there may ultimately be
some subtlety in the comparison of \Eq{\ref{moment}} and
\Eq{\ref{me}}.
For simplicity, in the following the renormalization scale $\mu$ will
be dropped and $\langle x\rangle^{u-d}$ will be understood as
evaluated in the $\overline{MS}$-scheme with $\mu=2~\mathrm{GeV}$.
\section{Persistent puzzle}
The puzzle with {$\langle x \rangle^{u-d}$} began with the earliest quenched lattice
calculations. As an example, in \Fig{\ref{quenched}}
\begin{figure}
\includegraphics[width=0.85\textwidth,angle=0]{quenched_x}
\label{qcdsfx}
\caption{Example quenched results for {$\langle x \rangle^{u-d}$}. The quenched results for
{$\langle x \rangle^{u-d}$} from~\cite{Gockeler:2002ek} are plotted against the square of
the pion mass $m_{PS}$ in units of the inverse Sommer scale
$r_0^{-1}$. The combination $(r_0 m_{PS})^2$ is proportional to the
quark mass in some units as the chiral limit is approached.
The quenched results for {$\langle x \rangle^{u-d}$} have a quite mild quark mass,
equivalently pion mass, dependence over a large range of pion
masses. Additionally, the linear extrapolation of this pion mass
dependence results in a substantial overestimate of the
phenomenologically determined value for {$\langle x \rangle^{u-d}$}. The quenched
approximation was a potential source of this discrepancy that has
since been eliminated. This plot was taken
from~\cite{Gockeler:2002ek}.}
\label{quenched}
\end{figure}
I show a quenched calculation of average-$x$
from~\cite{Gockeler:2002ek}. As seen there, {$\langle x \rangle^{u-d}$} has a mild, nearly
flat, pion mass dependence. Naively, this is not entirely unexpected.
Dimensionless quantities like {$\langle x \rangle^{u-d}$} tend to have a weaker dependence on
the quark mass than dimensionful quantities like the nucleon mass. In
fact, this sort of behavior would normally be welcomed, except in this
case the lattice calculation of {$\langle x \rangle^{u-d}$} clearly overshoots the
phenomenologically determined value.
At the time it was natural to dismiss this problem as simply being an
artifact of the quenched approximation that drops all contributions
from the so-called sea-quark loops. However, the results from the
earliest full QCD calculation~\cite{Dolgov:2002zm}, which included two
dynamical quark flavors, were found to agree with the quenched
calculations. The two-flavor results from~\cite{Dolgov:2002zm} are
shown in \Fig{\ref{chiral}},
\begin{figure}
\includegraphics[width=0.75\textwidth,angle=0]{chipt_x}
\caption{First full QCD results for {$\langle x \rangle^{u-d}$}. The results for {$\langle x \rangle^{u-d}$}
from~\cite{Dolgov:2002zm} are shown. They are plotted against the
square of the pion mass. Similar to the quenched results, average-$x$
from full QCD calculations, like the one shown here, have a mild
pion mass dependence and overshoot the phenomenological value for
{$\langle x \rangle^{u-d}$}. Chiral perturbation theory predicts the leading pion mass
dependence of {$\langle x \rangle^{u-d}$}~\cite{Arndt:2001ye,Chen:2001eg}. When combined
with a physically motivated regulator~\cite{Detmold:2001jb}, the
resulting functional form was capable of smoothly matching the
lattice computation to the expected pion mass dependence in the
chiral limit. Thus it was hypothesized that chiral dynamics might
help explain the seemingly strong quark mass dependence that would
be required to make the lattice results agree with the physical
value of {$\langle x \rangle^{u-d}$}. This explanation is being challenged by current
calculations. This plot was taken from~\cite{Dolgov:2002zm}.}
\label{chiral}
\end{figure}
where again the lattice calculations came out too high.
Staying with \Fig{\ref{chiral}}, another explanation for the puzzling
behavior of {$\langle x \rangle^{u-d}$} was put forth. The pion mass dependence of average-$x$
was calculated in chiral perturbation theory~\cite{Arndt:2001ye,
Chen:2001eg} and combined with a phenomenological
regulator~\cite{Detmold:2001jb} that was capable of accommodating both
the lattice calculation and the physical value of {$\langle x \rangle^{u-d}$}. It was
understood that the pion masses were too heavy to apply chiral
perturbation but~\cite{Detmold:2001jb} offered a plausibility argument
that chiral dynamics may lead to a strong quark mass dependence for
yet lighter quark masses while producing a mild quark mass dependence
for the range of quark masses used in contemporary lattice
computations. This was put on a slightly stronger footing with
calculations to higher-order in the chiral
expansion~\cite{Dorati:2007bk}. It was shown that appropriate choices
of the undetermined counterterms in the resulting functional form
could lead to a flat pion mass dependence for heavy pion
masses~\cite{Dorati:2007bk, Hagler:2007xi}.
This line of reasoning has dominated the lattice effort on nucleon
structure for much of the last decade. It was understood that
physically motivated regulators would introduce model dependence to
the extrapolation of the lattice calculations. It also seemed that
higher-order calculations would require the determination of too many
extra counterterms and low energy constants to be practically useful.
Thus the hope was to push to light enough pion masses to directly
observe the missing \emph{chiral logarithms}. These are the
contributions to {$\langle x \rangle^{u-d}$} of the form $m_\pi^2 \ln m_\pi^2$ that are, more
or less, uniquely predicted by chiral perturbation theory.
Fully dynamical lattice calculations of {$\langle x \rangle^{u-d}$} have continued to lighter
quark masses in search of these missing logarithms. The initial
two-flavor calculations~\cite{Dolgov:2002zm, Lin:2008uz, Baron:2008zz,
Pleiter:2011gw} have been extend to include the strange
quark~\cite{Ohta:2010sr, Hagler:2007xi, Winter:2011ze} and extended
further to even include the charm quark~\cite{Dinter:2011jt}. The
basic observation from the earliest quenched calculations remains
correct:\ average-$x$ appears to have a mild pion mass dependence, the
extrapolation of which is higher than the physical value. The
lightest pion mass used in the calculations referenced so far was
approximately $250~\mathrm{MeV}$. Understanding that finite-size
effects and lattice artifacts are seldom checked at the lightest pion
mass used in a calculation, we could argue that the lightest reliable
pion mass was closer to $300~\mathrm{MeV}$. This left some room for
the rapid pion mass dependence that would be required to reconcile the
lattice computation with the experimental measurement of {$\langle x \rangle^{u-d}$}, but
recent calculations have begun to challenge this scenario.
\section{Recent results}
Many of the most recent results were summarized
in~\cite{Alexandrou:2010cm}, but rather than showing all the
calculations of average-$x$, I focus on the results from the QCDSF
collaboration~\cite{Pleiter:2011gw}. Their calculation of {$\langle x \rangle^{u-d}$}
illustrates the most recent trend in lattice calculations:\ several
collaborations are now calculating at or near the physical pion
mass~\cite{Ohta:2011nv, Durr:2010aw, Aoki:2009ix}. There are a
variety of compromises that are made to accomplish this, but it is
still a very important advance. The calculation of {$\langle x \rangle^{u-d}$}
from~\cite{Pleiter:2011gw}, with pion masses approaching the physical
pion mass, is shown in \Fig{\ref{qcdsf}}.
\begin{figure}
\includegraphics[width=0.8\textwidth,angle=0]{x_qcdsf}
\caption{Recent results from the QCDSF collaboration for {$\langle x \rangle^{u-d}$}.
Preliminary results from the QCDSF
collaboration~\cite{Pleiter:2011gw} for {$\langle x \rangle^{u-d}$} are plotted versus the
square of the pion mass. These results rather dramatically continue
the long trend of lattice calculations of {$\langle x \rangle^{u-d}$} with quite mild pion
mass dependence that extrapolates to values noticeably higher than
the experimental measurement. These results are challenging the
prevailing view that chiral dynamics would cause sufficient
curvature in $m_\pi^2$ to reconcile the lattice calculations at
heavy pion masses with the value of {$\langle x \rangle^{u-d}$} at the physical point.
Also shown are a set of the most recent results for {$\langle x \rangle^{u-d}$} from global
analyses. These were collected in~\cite{Renner:2010ks} using
results from~\cite{Alekhin:2009ni, Blumlein:2006be, Blumlein:2004ip,
JimenezDelgado:2008hf, Martin:2009bu, Alekhin:2006zm}. Note that all these results are N${}^2$LO except
for the one explicitly marked as N${}^3$LO. The
results in this plot were communicated privately by the QCDSF
collaboration.}
\label{qcdsf}
\end{figure}
It is very plain to see that these latest results for average-$x$
confirm the nearly flat pion mass dependence of {$\langle x \rangle^{u-d}$} down to
essentially the physical point. This calculation achieves a long
sought milestone, but the conclusion is far from clear.
If the results from~\cite{Pleiter:2011gw} are taken at face value,
then it is hard to escape the obvious conclusion that one would draw
from \Fig{\ref{qcdsf}}. Either there are unaccounted for sources of
error in the lattice computation (or the global fits) or there is a
sizable discrepancy between the lattice determination of {$\langle x \rangle^{u-d}$} and the
experimental measurement of it. This later option seems unlikely, so
the current view among those doing the lattice calculations of {$\langle x \rangle^{u-d}$} is
that one or more of the systematic errors that must be checked for in
lattice calculations is currently underestimated.
Regarding the possibility of underestimated errors in the value of
{$\langle x \rangle^{u-d}$} extracted from the experimental measurements, I have shown the
results from six recent analyses of average-$x$ in \Fig{\ref{qcdsf}}.
There is some spread beyond that expected by the quoted errors on
{$\langle x \rangle^{u-d}$}, but it is certainly not large enough to account for the
difference between the lattice results and the global fits. Thus it
seems unlikely that there is a significant problem in the
phenomenological results for {$\langle x \rangle^{u-d}$}, but it is useful to keep in mind
that an extrapolation in $x$ is required to evaluate the integral in
\Eq{\ref{moment}}. Additionally, lattice calculations
often fail to specify the order to which the matching to $\overline{MS}$ is done,
but the difference between the experimental N${}^2$LO and N${}^3$LO results
in \Fig{\ref{qcdsf}}
suggests that this effect is small.
\section{Systematic errors}
The resolution of the puzzle in \Fig{\ref{qcdsf}} will likely hinge on
a careful examination of all the systematic errors present in the
lattice calculation of {$\langle x \rangle^{u-d}$}. Most of these sources of uncertainty
have been checked, to some extent, in previous calculations, so it was
believed that the dominant source of error in {$\langle x \rangle^{u-d}$} was due to the
poorly constrained extrapolation in the pion mass. However, the
results in \Fig{\ref{qcdsf}} now suggest that this might not be the
case. Certainly, the chiral extrapolation is no longer the single
stand out systematic error. This raises the possibility that other
errors were underestimated or that the discrepancy in
\Fig{\ref{qcdsf}} could be a combination of several smaller
uncertainties.
The advantage of using a renormalizable description of the fundamental
theory is that we know with confidence that the list of potential
errors is very limited. First, we have to reliably calculate the
basic nucleon matrix element in \Eq{\ref{me}}. This involves the
underlying algorithms used to stochastically evaluate the functional
integrals that define the matrix element. Since these methods are
used in many successful lattice computations, it seems unlikely that
there is a special algorithmic problem for the nucleon. Calculating
the matrix element in \Eq{\ref{me}} also requires isolating the ground
state, corresponding to the nucleon, using Euclidean space methods.
Several ongoing investigations~\cite{privQCDSF,privETMC} suggest that
this may be responsible for some, but not all, of the discrepancy in
{$\langle x \rangle^{u-d}$}. This can be called the \emph{plateau problem} because most
calculations of nucleon matrix elements rely on finding a plateau as a
function of Euclidean time in appropriately chosen correlation
functions. This issue has been examined off and on, most recently
in~\cite{Capitani:2010sg}, and a variety of new methods have been
developed to address it~\cite{Blossier:2009kd, Foley:2010vv}.
Once we have calculated the so-to-speak bare matrix element, the
operator renormalization must be accounted for. This is now regularly
calculated using nonperturbative methods, thus eliminating one
potential source of error. However, the method for renormalizing
composite operators nonperturbatively has its own set of potential
systematic errors. Most of these are quite technical in nature and go
well beyond the level of presentation in these proceedings, but the
overarching concern regards the separation of scales that is necessary
to nonperturbatively match the lattice operator to the continuum
$\overline{MS}$ operator needed in \Eq{\ref{me}}. Because
$\overline{MS}$ is a perturbative scheme, the matching must ultimately
involve some form of perturbation theory. To reduce the error from
this, the matching is done, ideally, at large renormalization scales
$\mu$, but this runs afoul of the constraint $\mu \ll 1/a$ that must
be maintained to control lattice cut-off effects.
There is some indirect evidence for a problem in the renormalization.
In~\cite{Renner:2010ks} it was pointed out that there does appear to
be some variation in the relative normalization of the results for
{$\langle x \rangle^{u-d}$} from different actions. Additionally, it was be found that
ratios of observables that cancel the renormalization lead to results
in agreement with experimental measurements~\cite{Pleiter:2011gw,
Ohta:2011nv}. Additionally, the renormalization is an
interesting potential culprit because it would lead to a simple
rescaling of {$\langle x \rangle^{u-d}$}. This is because $\overline{MS}$ is a mass
independent scheme, so the renormalization of the operator depends
only on the lattice spacing and not the quark masses. Thus a mistake
in the renormalization would correspond to a multiplicative rescaling
of \Fig{\ref{qcdsf}}, for example. It seems unlikely that such an
effect would account for the entire discrepancy, but it may be one
piece of the puzzle. As a separate cross check, there are attempts to
eliminate the renormalization issue entirely and calculate moments of
structure functions directly~\cite{Bietenholz:2010kn}.
Having reliably calculated the properly renormalized matrix element,
then the only remaining systematic errors are the extrapolation of the
heavier-than-physical pion mass to the physical pion mass and the
continuum and infinite-volume limits. Collectively, the results from
all the calculations of {$\langle x \rangle^{u-d}$} suggest that these errors are small
within the range of pion mass, lattice spacing and volumes that have
been used. The concern, though, is that the asymptotic values of each
of the three limits may not have been seen in the currently used
ranges. The size of lattice artifacts can be checked by establishing
not just weak lattice spacing dependence, as is customary, but by
demonstrating the expected scaling as the continuum limit is
approached. For the $L$ dependence, one could explicitly check the
generically expected exponential suppression at large $L$, rather than
the current standard of simply demonstrating an apparent convergence
to within the errors. It is hard to point to any compelling
indications of finite-size or cut-off effects in the current
calculations of {$\langle x \rangle^{u-d}$}, but that may be so because these issues have
never been pursued with the high precision likely required to detect
such effects.
The pion mass dependence is harder to check because the only
expectations come from chiral perturbation theory and that may simply
not be applicable for the physical pion mass or heavier. But as
\Fig{\ref{qcdsf}} demonstrates, current calculations are quickly
reducing the extent of the required extrapolation in the pion mass and
not-so-far-off calculations will be able to bridge the physical pion
mass, thus converting an extrapolation into an interpolation.
We also must keep in mind that each of these possible systematics can
interfere with each other. For example, failing to reliably determine
the matrix element may produce results that erroneously have a flat
pion mass dependence or failing to properly renormalize the needed
operator may obscure the cut-off dependence. And of course, all the
calculations must ultimately be done with sufficient statistical
precision to be capable of clearly checking the, relatively short,
list of systematic errors.
\section{Conclusions}
Lattice calculations are proceeding steadily down to the physical pion
mass. This has facilitated the precision calculation of many QCD
observables, however, it has also produced some puzzles. In
particular, recent calculations of average-$x$ have continued a long
established trend of overshooting the value for {$\langle x \rangle^{u-d}$} determined by
global analyses. For quite some time, this had been assumed to be
caused by a suppression of chiral dynamics due to the use of
heavier-than-physical pion masses, but this explanation is much less
compelling in the light of recent results. Clarification of this
situation will likely require a careful study of all the possible
uncertainties in the computation of {$\langle x \rangle^{u-d}$}. The use of a well-defined
nonperturbative regulator, namely lattice QCD, ensures that the
errors in the calculation of {$\langle x \rangle^{u-d}$} are identifiable and all
systematically improvable. Controlling the chiral limit is still, of
course, one source of error in the lattice calculations, but it is now
just one among several potentially comparable errors. This marks a
milestone in the lattice effort to determine hadron structure directly
from QCD.
\begin{theacknowledgments}
This work was supported by SciDAC and by Jefferson Science Associates,
LLC under U.S. DOE Contract No. DE-AC05-06OR23177. The U.S. Government
retains a non-exclusive, paid-up, irrevocable, world-wide license to
publish or reproduce this manuscript for U.S. Government purposes.
\end{theacknowledgments}
\bibliographystyle{aipproc}
|
\section{Introduction}
One of the fundamental fields of research in infinite combinatorics is the study of partition relations.
To recall some results in this area, we shall need the following piece of notation.
For cardinals $\kappa$ and $\theta$, the (negative) partition relation $\kappa\not\rightarrow[\kappa]^2_\theta$
stands for the existence of a function $f:[\kappa]^2\rightarrow\theta$ such that
$f``[A]^2=\theta$ for every $A\s \kappa$ of size $\kappa$.
In a paper from 1929, Ramsey proved
that $\omega\not\rightarrow[\omega]^2_2$ \emph{fails},
that is, every symmetric function $f:[\omega]^2\rightarrow 2$ admits an infinite homogenous set.
A few years later, Sierpi\'nski
proved that Ramsey's theorem does not generalize to higher cardinals, introducing
a function witnessing $\omega_1\not\rightarrow[\omega_1]^2_2$.
At the 1960's,
Erd\"os, Hajnal and Rado \cite{partitionrelationsforcardinalnumbers} proved that the Generalized Continuum Hypothesis ($\gch$) implies
$\lambda^+\not\rightarrow[\lambda^+]^{2}_{\lambda^+}$ for every infinite cardinal $\lambda$ (see also \cite[Theorem 49.1]{combintorialsettheory}),
and at the mid 1980's, Todorcevic \cite{todorcevic} made a breakthrough, proving outright in $\zfc$ that $\lambda^+\not\rightarrow[\lambda^+]^2_{\lambda^+}$
holds for every regular cardinal $\lambda$.
The question of whether $\lambda^+\not\rightarrow[\lambda^+]^2_{\lambda^+}$ may fail for a singular cardinal $\lambda$ is still open.
In parallel, throughout the years, a study of stronger notions of negative partition relations was carried (see, for instance, Shelah \cite{sh276},\cite{sh:g},\cite{sh572},
and Shelah-Eisworth \cite{sh535},\cite{EiSh:819}).
In this paper, we shall be interested in the following two particular strengthenings:
\begin{itemize}
\item (Rectangular form) $\kappa\not\rightarrow[\kappa;\kappa]^2_\theta$
asserts the existence of a function $f:[\kappa]^2\rightarrow\theta$ such that
$f[A\times B]=\theta$ for every $A\s\kappa,B\s\kappa$ of size $\kappa$.
\item (Shelah's strong coloring \cite{sh:g}) $\pr_1(\kappa,\kappa,\theta,\sigma)$ asserts the existence of a function $f:[\kappa]^2\rightarrow\theta$ such that
for every $\mathcal A\subseteq\mathcal P(\kappa)$ of size $\kappa$, consisting of pairwise disjoint sets of size $<\sigma$, and every $\xi<\theta$,
there exists distinct $a,b\in\mathcal A$ such that $f[a\times b]=\{\xi\}$.
\end{itemize}
Comparing the above three concepts is, of course, a very natural task.
To appreciate this task,
we mention that while the following results goes back to the work of Todorcevic from the 1980's:
\begin{itemize}
\item $\omega_1\not\rightarrow[\omega_1]^2_{\omega_1}$;
\item $\pp(\lambda)=\lambda^+$
implies $\lambda^+\not\rightarrow[\lambda^+]^2_{\lambda^+}$ for every singular cardinal $\lambda$,
\end{itemize}
the stronger versions was proved rather recently:
\begin{itemize}
\item (Moore, 2006 \cite{lspace}) $\omega_1\not\rightarrow[\omega_1;\omega_1]^2_{\omega_1}$;
\item (Eisworth, 2009 \cite{eisworth2}) $\pp(\lambda)=\lambda^+$ implies $\pr_1(\lambda^+,\lambda^+,\lambda^+,\cf(\lambda))$
for every singular cardinal $\lambda$.
\end{itemize}
In \cite{rinot12}, the author introduced the \emph{Ostaszewski square} $\sc_\kappa$,
which is a generalization of Jensen's square principle $\square_\kappa$ from \cite{jensen1972}.
One of the application appearing in \cite{rinot12} addresses negative partition relations that are derived from Ostaszewski squares.
Of course, we were hoping to yield as stronger partition relations as possible,
and incidently, around the same time, Eisworth \cite{eisworth2} made a significant progress on the old question of comparing the different notions,
showing that some strong coloring statements
are indeed equivalent to seemingly weaker statements.
After studying Eisworth's works \cite{eisworth2},\cite{eisworth1}, we realized that his arguments may
be pushed further to completely clarify the situation at the successor of singular cardinals,
and indeed, the main result of this paper reads as follows.
\begin{THM}\label{t11} For every singular cardinal $\lambda$, and every $\theta\le\lambda^+$, the following are equivalent:
\begin{enumerate}
\item $\lambda^+\not\rightarrow[\lambda^+]^2_\theta$;
\item $\lambda^+\not\rightarrow[\lambda^+;\lambda^+]^2_\theta$;
\item $\pr_1(\lambda^+,\lambda^+,\theta,\cf(\lambda))$.
\end{enumerate}
\end{THM}
We mention that the above yields an alternative proof to Theorem 5 of \cite{eisworth2} (that is, a proof that does not appeal to \cite{sh572}),
as well as an affirmative answer to the question
of Eisworth raised in \cite{eisworth3} of
whether the failure of $\pr_1(\lambda^+,\lambda^+,\lambda^+,\cf(\lambda))$
implies that any collection of less than $\cf(\lambda)$ many stationary subsets of $\lambda^+$ must reflect simultaneously.
The actual proof of Theorem \ref{t11} goes through establishing the existence of
a function of its own interest: a function that transforms rectangles $A\circledast B$ into squares $[C]^2$.\footnote{Here,
$A\circledast B$ stands for the set $\{(\alpha,\beta)\in A\times B\mid \alpha<\beta\}$, and $[C]^2:=C\circledast C$.}
\begin{THM}\label{t12} For every singular cardinal $\lambda$,
there exists a function $\rts:[\lambda^+]^2\rightarrow[\lambda^+]^2$ such that
for every cofinal subsets $A,B$ of $\lambda^+$, there exists a cofinal $C\s\lambda^+$ such that $\rts[A\circledast B]\supseteq [C]^2$.
Moreover:
\begin{itemize}
\item for every cofinal subsets $A,B$ of $\lambda^+$, and every $\kappa<\lambda$,
there exists a stationary subset $S\s E^{\lambda^+}_{>\kappa}$ such that $\rts[A\circledast B]\supseteq [S]^2$;\footnote{Here, $E^{\lambda^+}_{>\kappa}$ stands for the set $\{\alpha<\lambda^+\mid \cf(\alpha)>\kappa\}$.}
\item if $\alpha<\beta<\lambda^+$ and $\rts(\alpha,\beta)=(\alpha^*,\beta^*)$, then $\alpha^*\le\alpha<\beta^*\le\beta$.
\end{itemize}
\end{THM}
\subsection*{Acknowledgement}
We would like to express our intellectual debt to Eisworth, Shelah, and Todorcevic,
whom technology is in the heart of this paper.
While the results of this paper yields an answer to a question of Eisworth from \cite{eisworth3},
the proof of $(2)\Rightarrow (3)$ in Theorem \ref{t11} builds on the main result from Eisworth's \cite{eisworth2}
and on Theorem \ref{t12}, whereas, the proof of Theorem \ref{t12} is just the outcome of several complications of Eisworth's arguments from \cite{eisworth1}.
\section{Transforming rectangles into squares}
In this section we prove Theorem \ref{t12}, where the proof splits into two cases --- the case $\cf(\lambda)>\omega$,
and the case $\cf(\lambda)=\omega$. In either case, we first define the function $\rts$, and then verify that it works.
We mention that the definition(s) of $\rts$ are virtually the same as that of the function $D$ from \cite[Theorem 1]{eisworth1},
and that our analysis of these functions is just a slight extension of Eisworth's analysis from \cite{eisworth1}.
\subsection{The uncountable cofinality case}
In this subsection, we prove Theorem \ref{t12}, for the case that $\lambda$ is a singular cardinal of uncountable cofinality.
We shall rely on Shelah's theorem on club guessing.
\begin{thm}[Shelah, \cite{sh:g}]\label{31} For every cardinal $\lambda>\cf(\lambda)>\omega$,
there exists a sequence $\overrightarrow e=\langle e_\delta\mid \delta\in
E^{\lambda^+}_{\cf(\lambda)}\rangle$ such that:
\begin{enumerate}
\item $e_\delta$ is a club in $\delta$ of order-type $\cf(\lambda)$, for all $\delta\in E^{\lambda^+}_{\cf(\lambda)}$;
\item for every club $E\s\lambda^+$, there exist stationarily many $\delta\in E^{\lambda^+}_{\cf(\lambda)}$ for which
$\sup\{ \cf(\alpha)\mid \alpha\in e_\delta\cap E\}=\lambda$.
\end{enumerate}
\end{thm}
\begin{lemma}[Shelah, \cite{sh:g}]\label{32} Let $\lambda$ and $\overrightarrow e$ be as in the preceding theorem.
Then there exists a sequence $\overrightarrow C=\langle C_\alpha\mid
\alpha<\lambda^+\rangle$ such that for every $\alpha<\lambda^+$:
\begin{itemize}
\item $\max(C_{\alpha+1})=\alpha$;
\item $C_\alpha$ is a club in $\alpha$ of size $<\lambda$;
\item if $\delta\in C_\alpha\cap E^{\lambda^+}_{\cf(\lambda)}$, then $e_\delta\s C_\alpha$.
\end{itemize}
\end{lemma}
From now on, we fix a cardinal $\lambda$ and a sequence
$\overrightarrow e,\overrightarrow C$ as in the preceding, and shall be conducting \emph{minimal walks} \cite{MR2355670}
along this $\overrightarrow C$-sequence.
For $\alpha<\beta<\lambda^+$, let $\beta_0:=\beta$, and $\beta_{n+1}:=\min(C_{\beta_n}\bks\alpha)$
whenever $n<\omega$ and $\beta_n>\alpha$.
The decreasing sequence $\beta=\beta_0>\beta_1>\ldots\beta_{k}=\alpha$ is called the
$\overrightarrow C$-walk from $\beta$ down to $\alpha$, and we denote
\begin{itemize}
\item $\rho(\alpha,\beta):=k$;
\item $\Tr(\alpha,\beta):=\{ \beta_0,\ldots,\beta_k\}$;
\item $\tr(\alpha,\beta):=\Tr(\alpha,\beta)\bks\{\alpha\}$.
\end{itemize}
Next, by Shelah \cite{sh:g}, we may fix a scale $(\overrightarrow\lambda,\overrightarrow f)$ for $\lambda$.
That is, a sequence $\overrightarrow\lambda=\langle \lambda_i\mid i<\cf(\lambda)\rangle$
of regular cardinals which is strictly increasing and cofinal in $\lambda$,
and a sequence $\overrightarrow f=\langle f_\alpha\mid\alpha<\lambda^+\rangle$
of elements from $\prod\overrightarrow\lambda$ which is strictly increasing and cofinal in $\left\langle \prod\overrightarrow\lambda,\le^*\right\rangle$.
So, $\alpha<\beta<\lambda^+$ implies that $\{ i<\cf(\lambda)\mid f_\alpha(i)\ge f_\beta(i)\}$ is of size $<\cf(\lambda)$.
In particular, it makes sense to consider the next ordinal:
$$\Delta(\alpha,\beta):=\sup\{ i<\cf(\lambda)\mid f_\alpha(i)\ge f_\beta(i)\}.$$
Let $\psi:\cf(\lambda)\rightarrow\omega\times\omega$ be some
surjection such that the pre-image of every element is cofinal in $\cf(\lambda)$.
Fix $\alpha<\beta<\lambda^+$, and let us define $\rts(\alpha,\beta)$.
Let $k_1$, $k_2$ be such that
$\psi(\Delta(\alpha,\beta))=(k_1,k_2).$
Next, let:
\begin{itemize}
\item $\beta_0>\dots>\beta_{n}$ denote the $\overrightarrow C$-walk from $\beta$ down to $\alpha$;
\item $\eta:=\max\{\sup(C_{\tau}\cap\beta_{k_1})\mid \tau\in\tr(\beta_{k_1},\beta)\}$;
\item $\alpha_0>\ldots>\alpha_{m}$ denote the $\overrightarrow C$-walk from $\alpha$ down to $\eta+1$;
\end{itemize}
We would like to assign $\rts(\alpha,\beta):=(\alpha_{k_2},\beta_{k_1})$.
Of course, if $k_1> n$, $\alpha<\eta+1$, or $k_2> m$,
then the above definition would not make sense. In such cases, we just put $\rts(\alpha,\beta):=(\alpha,\beta)$.
In Theorem \ref{t25} below, we prove that $\rts$ has the required transfer property.
The proof builds on a technical lemma (Lemma \ref{344}), whose statement requires the following piece of notation.
\begin{defn}\label{dddd} For $A\s\lambda^+$, $\eta\le\epsilon<\lambda^+$, and $k<\omega$, let
$$A_{k,\eta}(\epsilon):=\left\{\alpha\in A\mid \rho(\epsilon,\alpha)=k, \max\{\sup(C_\tau\cap\epsilon)\mid \tau\in\tr(\epsilon,\alpha)\}=\eta\right\}.$$
\end{defn}
\begin{lemma}\label{344} If $A,B$ are cofinal subsets of $\lambda^+$, and $\theta<\lambda$,
then the following set is stationary:
$$S^{A,B}_{>\theta}:=\{\epsilon\in E^{\lambda^+}_{>\theta}\mid \exists(k_1,k_2,\eta)\in\omega\times\omega\times\epsilon\left( |B_{k_1,\eta}(\epsilon)|=|A_{k_2,\eta}(\epsilon)|=\lambda^+\right)\}.$$
\end{lemma}
\begin{proof} Suppose that $A,B$ are cofinal subsets of $\lambda^+$, $\theta<\lambda$,
and that $E$ is a club subset of $\lambda^+$.
We shall prove that $S^{A,B}_{>\theta}\cap E\not=\emptyset$.
Let $S$ denote the set of all $\delta\in E^{\lambda^+}_{\cf(\lambda)}$ for which there exists
an elementary submodel $M\prec\mathcal H(\chi)$ such that:\footnote{As usual, $\chi$ is some large enough regular cardinal which we fix throughout the whole paper,
$\mathcal H(\chi):=\langle H(\chi),\in,<_\chi\rangle$ is an hereditary fragment of the universe that admits a well-ordering $<_\chi$.}
\begin{itemize}
\item $\{\overrightarrow C,A,B\}\s M$;
\item $M\cap\lambda^+=\delta>\lambda$;
\item $\sup\{ \cf(\epsilon)\mid \epsilon\in e_\delta\cap E\}=\lambda$.
\end{itemize}
By the choice of the club guessing sequence $\overrightarrow e$, the set $S$ is stationary.
Put $D:=\{\varepsilon<\lambda^+\mid \sup(S\cap\varepsilon)=\varepsilon\}$.
Then $D$ is a club,
so let us fix some $\nu\in E^{\lambda^+}_{\cf(\lambda)}$ such that
$\sup\{ \cf(\varepsilon)\mid \varepsilon\in e_\nu\cap D\}=\lambda$.
Put $\beta:=\min(B\bks\nu+1)$, and $\alpha:=\min(A\bks\nu+1)$.
Let
\begin{itemize}
\item $\beta=\beta_0>\ldots>\beta_{n+1}=\nu$ denote the $\overrightarrow C$-walk from $\beta$ down to $\nu$;
\item $\alpha=\alpha_0>\ldots>\alpha_{m+1}=\nu$ denote the $\overrightarrow C$-walk from $\alpha$ down to $\nu$;
\item $\gamma:=\max(\{\sup(C_{\beta_i}\cap\nu)\mid i<n\}\cup \{\sup(C_{\alpha_j}\cap\nu)\mid j<m\}).$
\end{itemize}
Since $\nu\not\in C_{\beta_i}\cup C_{\alpha_j}$ for all $i<n$ and $j<m$,
we get that $\gamma<\nu$, so let us pick some large enough
$\varepsilon\in e_\nu\cap D$ above $\gamma$ such that
$\cf(\varepsilon)>\max\{|C_{\beta_n}|,|C_{\alpha_m}|\}$.
\begin{claim}\label{c241} $\tr(\varepsilon,\beta)=\tr(\nu,\beta)$, and $\tr(\varepsilon,\alpha)=\tr(\nu,\alpha)$.
\end{claim}
\begin{proof}
As $(\gamma,\nu)\cap C_{\beta_i}=\emptyset$
for all $i<n$, and $\varepsilon\in(\gamma,\nu)$, we get that
$\min(C_{\beta_i}\bks\varepsilon)=\min(C_{\beta_i}\bks\nu)$ for all $i<n$,
and hence $\Tr(\varepsilon,\beta)\bks\beta_n=\{\beta_i\mid i\le n\}=\tr(\nu,\beta)$.
Likewise, by $(\gamma,\nu)\cap C_{\alpha_j}=\emptyset$
for all $j<m$, we get that
$\Tr(\varepsilon,\alpha)\bks\alpha_m=\tr(\nu,\alpha)$.
Since $\nu\in
C_{\beta_n}\cap C_{\alpha_m}\cap E^{\lambda^+}_{\cf(\lambda)}$, we have $e_\nu\s
C_{\beta_n}\cap C_{\alpha_m}$, and in particular $\varepsilon\in C_{\beta_n}\cap C_{\alpha_m}$. So
$\rho(\varepsilon,\beta)=n+1$, $\rho(\varepsilon,\alpha)=m+1$,
and hence
$\tr(\varepsilon,\beta)=\tr(\nu,\beta)$, $\tr(\varepsilon,\alpha)=\tr(\nu,\alpha)$.
\end{proof}
Put $$\gamma':=\max\{\sup(C_{\tau}\cap\varepsilon)\mid
\tau\in\tr(\varepsilon,\beta)\cup\tr(\varepsilon,\alpha)\}.$$
Since $\gamma<\varepsilon<\nu$, we have $\gamma'=\max\{\gamma,\sup(C_{\beta_n}\cap\varepsilon),\sup(C_{\alpha_m}\cap\varepsilon)\}$.
Since $\varepsilon\in C_{\beta_n}\cap C_{\alpha_m}$ and $\cf(\varepsilon)>\max\{|C_{\beta_n}|,|C_{\alpha_m}|\}$, we
get that $\max\{\sup(C_{\beta_n}\cap\varepsilon),\sup(C_{\alpha_m}\cap\varepsilon)\}<\varepsilon$,
and hence $\gamma'<\varepsilon$.
Since $\varepsilon\in e_\nu\cap D$, we now pick some large enough $\delta\in S$ satisfying $\gamma'<\delta<\varepsilon$,
and consider additional walks. Let:
\begin{itemize}
\item $\beta=\beta'_0>\ldots>\beta'_{n'+1}=\delta$ denote the $\overrightarrow C$-walk
from $\beta$ down to $\delta$;
\item $\alpha=\alpha'_0>\ldots>\alpha'_{m'+1}=\delta$ denote the $\overrightarrow C$-walk from $\alpha$ down to $\delta$.
\end{itemize}
\begin{claim} \begin{enumerate}
\item $\tr(\delta,\beta)=\tr(\delta,\varepsilon)\cup\tr(\varepsilon,\beta)$;
\item $\tr(\delta,\alpha)=\tr(\delta,\varepsilon)\cup\tr(\varepsilon,\alpha)$.
\end{enumerate}
\end{claim}
\begin{proof}
By definition of $\gamma'$ and since $\delta\in(\gamma',\varepsilon)$, we get that $\beta'_i=\beta_i$ for all $i\le n$,
and hence $\beta'_{n+1}=\varepsilon$. So $\tr(\varepsilon,\beta)=\{\beta'_0,\ldots\beta'_n\}$, and $\tr(\delta,\varepsilon)=\{\beta'_{n+1},\ldots,\beta'_{n'}\}$.
Likewise, $\alpha'_j=\alpha_j$ for all $j\le m$, and hence $\alpha'_{m+1}=\varepsilon$.
So $\tr(\varepsilon,\alpha)=\{\alpha'_0,\ldots,\alpha'_m\}$,
and $\tr(\delta,\varepsilon)=\{\alpha'_{m+1},\ldots,\alpha'_{m'}\}$.
\end{proof}
Next, denote
\begin{itemize}
\item $\zeta:=\min(\tr(\delta,\varepsilon))$;
\item $\gamma_\alpha:=\max\{\sup(C_{\alpha'_j}\cap\delta)\mid j<m'\}$;
\item $\gamma_\beta:=\max\{\sup(C_{\beta'_i}\cap\delta)\mid i<n'\}$.
\end{itemize}
Evidently, $\gamma_\alpha<\delta$, $\gamma_\beta<\delta$. Note that by the preceding claim, we have $\beta'_{n'}=\alpha'_{m'}=\zeta$.
Pick a large enough $\epsilon\in e_\delta\cap E$ such that
$\sup(e_\delta\cap\epsilon)>\max\{\gamma_\alpha,\gamma_\beta\}$, and
$\cf(\epsilon)>\max\{|C_{\zeta}|,\theta\}$.
\begin{claim} $\tr(\epsilon,\beta)=\tr(\delta,\beta)$, and $\tr(\epsilon,\alpha)=\tr(\delta,\alpha)$.
\end{claim}
\begin{proof} As in the proof of Claim \ref{c241}.
\end{proof}
Denote:
\begin{itemize}
\item $\eta_\alpha:=\max\{\sup(C_\tau\cap\epsilon)\mid \tau\in\tr(\epsilon,\alpha)\}$;
\item $\eta_\beta:=\max\{\sup(C_\tau\cap\epsilon)\mid \tau\in\tr(\epsilon,\beta)\}$.
\end{itemize}
Since $\min(\tr(\epsilon,\beta))=\min(\tr(\delta,\beta))=\min(\tr(\delta,\varepsilon))=\zeta$,
and $\gamma_\beta<\epsilon<\delta$, we get that $\eta_\beta=\max\{\gamma_\beta,\sup(C_\zeta\cap\epsilon)\}$.
Since $\delta\in C_\zeta\cap E^{\lambda^+}_{\cf(\lambda)}$, we get that $\sup(C_\zeta\cap\epsilon)\ge\sup(e_\delta\cap\epsilon)>\gamma_\beta$,
and hence $\eta_\beta=\sup(C_\zeta\cap\epsilon)$.
By a similar consideration, we get that $\eta_\alpha=\sup(C_\zeta\cap\epsilon)$.
Denote $\eta:=\sup(C_\zeta\cap\epsilon)$.
Then, we have just established that
\begin{itemize}
\item $\beta\in B_{k_1,\eta}(\epsilon)$, for $k_1:=\rho(\epsilon,\beta)$, and
\item $\alpha\in A_{k_2,\eta}(\epsilon)$, for $k_2:=\rho(\epsilon,\alpha)$.
\end{itemize}
By the choice of $\delta$, let us now pick an elementary submodel $M\prec\mathcal H(\chi)$
such that $\{A,B,\overrightarrow C\}\s M$ and $M\cap\lambda^+=\delta$.
By $\eta\le\epsilon<\delta$,
the sets $B_{k_1,\eta}(\epsilon)$ and $A_{k_2,\eta}(\epsilon)$ are definable from parameters in $M$,
and hence belongs to $M$. Since $|M|=\lambda$ and $\beta\in B_{k_1,\eta}(\epsilon)\bks M$,
we get that $|B_{k_1,\eta}(\epsilon)|=\lambda^+$.
Likewise, $|A_{k_2,\eta}(\epsilon)|=\lambda^+$.
By the choice of $\epsilon$,
we have $\cf(\epsilon)>|C_\zeta|$, and hence $\eta<\epsilon$.
Altogether, $\epsilon\in S^{A,B}_{>\theta}$.
Recalling that $\epsilon\in e_\delta\cap E$, we conclude that
$S^{A,B}_{>\theta}\cap E\not=\emptyset$.
\end{proof}
\begin{thm}\label{t25} For every cofinal subsets $A,B$ of $\lambda^+$, and every $\theta<\lambda$,
there exists a stationary subset $S^*\s E^{\lambda^+}_{>\theta}$ such that $\rts[A\circledast B]\supseteq [S^*]^2$.
\end{thm}
\begin{proof}
Suppose that $A,B$ are cofinal subsets of $\lambda^+$, and that $\theta<\lambda$.
By Lemma \ref{344}, $S^{A,B}_{>\theta}$ is stationary, so we appeal to Fodor's
lemma and fix some $k_1,k_2<\omega$ and $\eta<\lambda^+$ such that the
following set is stationary:
$$S:=\{\epsilon\in E^{\lambda^+}_{>\theta}\mid |B_{k_1,\eta}(\epsilon)|=|A_{k_2,\eta}(\epsilon)|=\lambda^+\}.$$
Next, let $\langle M_i\mid i<\lambda^+\rangle$ be a continuous
$\in$-chain of elementary submodels of $\mathcal H(\chi)$ such that:
\begin{enumerate}
\item $\{\overrightarrow \lambda,\overrightarrow f,\overrightarrow C,A,\eta\}\s M_0$;
\item $|M_i|=\lambda$ for all $i<\lambda$;
\end{enumerate}
Let $S^*$ be the following stationary subset of $E^{\lambda^+}_{>\theta}$:
$$S^*:=\{ \epsilon\in S\mid M_\epsilon\cap\lambda^+=\epsilon\}.$$
\begin{claim}\label{c251} For every $\alpha^*<\beta^*$ from $S^*$,
there exists $\alpha\in A$, $\beta\in B$ such that all of the followings holds:
\begin{enumerate}
\item $\eta<\alpha^*<\alpha<\beta^*<\beta$;
\item $\rho(\beta^*,\beta)=k_1$;
\item $\rho(\alpha^*,\alpha)=k_2$;
\item $\psi(\Delta(\alpha,\beta))=(k_1,k_2)$;
\item $\max\{\sup(C_\tau\cap\beta^*)\mid \tau\in \tr(\beta^*,\beta)\}=\eta$;
\item $\max\{\sup(C_\tau\cap\alpha^*)\mid \tau\in \tr(\alpha^*,\alpha)\}=\eta$;
\end{enumerate}
\end{claim}
\begin{proof}
As $\beta^*\in S$, let us pick some $\beta\in B_{k_1,\eta}(\beta^*)$ with $\beta^*<\beta$.
Next, let $M=\sk_{\mathcal H(\chi)}(x)$ denote the Skolem hull in $\mathcal H(\chi)$ of the set $x:=\{(\overrightarrow\lambda,\overrightarrow f),\overrightarrow
C,A,\eta,\alpha^*\}\cup\cf(\lambda)+1$. As $|M|<\lambda=\sup(\overrightarrow\lambda)$,
it is now reasonable to consider the characteristic function of $M$ on $\overrightarrow\lambda$.
We remind the reader that the latter, which we denote by $\ch^{\overrightarrow\lambda}_M$, stands for the unique
element of $\prod\overrightarrow\lambda$ that satisfies for all $i<\cf(\lambda)$:
$$\ch^{\overrightarrow\lambda}_M(i)=\begin{cases}\sup(M\cap\lambda_i),&\sup(M\cap\lambda_i)<\lambda_i\\
0,&\text{otherwise}\end{cases}.$$
Next, note that $M\in M_{\beta^*}$.
Indeed, as $x\s M_{\alpha^*+1}\prec\mathcal H(\chi)$, we get that
$\sk_{\mathcal H(\chi)}(x)=\sk_{M_{\alpha^*+1}}(x)$. So, $M$ is definable from $x$ and $M_{\alpha^*+1}$,
and hence $M\in M_{\alpha^*+2}\s M_{\beta^*}$.
By $M\in M_{\beta^*}$, we get that $\ch^{\overrightarrow\lambda}_M\in M_{\beta^*}$, and hence
$$M_{\beta^*}\models \exists\delta<\lambda^+(\ch^{\overrightarrow\lambda}_M<^*f_\delta).$$
As $M_{\beta^*}\cap\lambda^+=\beta^*$, we may find some
$\delta<\beta^*$ such that $\ch^{\overrightarrow\lambda}_M<^*f_\delta$.
By $\delta<\beta^*<\beta$, we infer that $\ch^{\overrightarrow\lambda}_M<^*f_\beta$.
Denote $A':=A_{k_2,\eta}(\alpha^*)\bks(\alpha^*+1)$.
For all $i<\cf(\lambda)$, put $\Gamma_i:=\{f_\alpha(i)\mid
\alpha\in A'\}$. As $A'$ is cofinal in $\lambda^+$, $\langle
f_\alpha\mid \alpha\in A'\rangle$ is cofinal in
$\langle\prod\overrightarrow\lambda,<^*\rangle$, and hence
$\sup(\Gamma_i)=\lambda_i$ for co-boundedly many $i<\cf(\lambda)$.
Thus, let us pick some large enough ${i'}<\cf(\lambda)$ such that:
\begin{itemize}
\item $\ch^{\overrightarrow\lambda}_M(i)<f_\beta(i)$ whenever $i'<i<\cf(\lambda)$;
\item $\sup(\Gamma_{{i'}})=\lambda_{{i'}}>|M|$.
\item $\psi(i')=(k_1,k_2)$.
\end{itemize}
Put $\varepsilon:=f_\beta(i')$. As $\varepsilon\in\lambda_{i'}$,
we define $N:=\sk_{\mathcal H(\chi)}(M\cup\lambda_{i'})$.
By $x\s M\s N$, we have
$\{\overrightarrow C,A,k_2,\eta,\alpha^*,\varepsilon\}\s N$.
So $A'\in N$, and
$$N\models \exists\alpha\in A'\left(f_\alpha(i')>\varepsilon\right).$$
Thus, let us pick some $\alpha\in A'\cap N$ such that
$f_\alpha(i')>\varepsilon$. Since $\alpha\in N\s M_{\beta^*}$, we
have $\alpha<\beta^*$.
So $\alpha\in A_{k_2,\eta}(\alpha^*)$, $\beta\in B_{k_1,\eta}(\beta^*)$,
$\eta<\alpha^*<\alpha<\beta^*<\beta$, and hence we are left with verifying that $\psi(\Delta(\alpha,\beta))=(k_1,k_2)$.
\begin{sclaim}\label{c251} Let $\ch_N^{\overrightarrow \lambda}$ denote the characteristic function of $N$ on $\overrightarrow\lambda$.
Then $\ch^{\overrightarrow\lambda}_N(i)=\ch^{\overrightarrow\lambda}_M(i)$ whenever $i'<i<\cf(\lambda)$.
\end{sclaim}
\begin{proof} This is an instance of Baumgartner's lemma \cite{baum},
stating that if $M\preceq\mathcal H(\chi)$ is some structure, and $\kappa<\mu$ are regular cardinals in $M$,
then $\sup(N\cap\mu)=\sup(M\cap\mu)$, for $N:=\sk_{\mathcal H(\chi)}(M\cup\kappa)$.
\end{proof}
Next, note that for all $i$ with $i'<i<\cf(\lambda)$, we have:
\begin{itemize}
\item $f_\alpha(i)\le\ch^{\overrightarrow\lambda}_N(i)$, since $\alpha\in N$;
\item $\ch^{\overrightarrow\lambda}_N(i)=\ch^{\overrightarrow\lambda}_M(i)$, by the preceding subclaim;
\item $\ch^{\overrightarrow\lambda}_M(i)<f_\beta(i)$, by $i>i'$ and the choice of $i'$.
\end{itemize}
So $f_\alpha(i)<f_\beta(i)$ whenever $i<i'<\cf(\lambda)$, and hence $\Delta(\alpha,\beta)\le i'$.
By $f_\alpha(i')>\varepsilon=f_\beta(i')$, we also get that $\Delta(\alpha,\beta)\ge i'$.
It follows that $\Delta(\alpha,\beta)= i'$, and hence $\psi(\Delta(\alpha,\beta))=\psi(i')=(k_1,k_2)$.
\end{proof}
\begin{claim} $[S^*]^2\s\rts``[A\otimes B]$.
\end{claim}
\begin{proof}
Suppose that $(\alpha^*,\beta^*)$ is a given element of
$[S^*]^2$. Fix $\alpha\in A$, $\beta\in B$
as in Claim \ref{c251}, and let us show that
$\rts(\alpha,\beta)=(\alpha^*,\beta^*)$.
Let:
\begin{itemize}
\item $\beta_0>\ldots>\beta_{n}$ denote the
$\overrightarrow C$-walk from $\beta$ down to $\alpha$;
\item $\beta'_0>\ldots>\beta'_{k_1}$ denote the
$\overrightarrow C$-walk from $\beta$ down to $\beta^*$;
\item $\alpha_0>\ldots>\alpha_{m}$ denote the
$\overrightarrow C$-walk from $\alpha$ down to $\eta+1$;\footnote{Notice that by the choice of $\beta$,
we have $\eta<\alpha^*<\alpha$, and hence $\alpha>\eta+1$.}
\item $\alpha'_0>\ldots>\alpha'_{k_2}$ denote the
$\overrightarrow C$-walk from $\alpha$ down to $\alpha^*$.
\end{itemize}
By $\beta_0'=\beta_0=\beta$ and $\sup(C_{\beta'_i}\cap\beta^*)\le \eta<\alpha<\beta^*$ for all $i<k_1$,
we get that $\beta_{i+1}'=\min(C_{\beta_{i'}}\bks \beta^*)=\min(C_{\beta_i}\bks\alpha)=\beta_{i+1}$ for all $i<k_1$.
In particular, $n\ge k_1$, $\beta_{k_1}=\beta^*$, and
$$\eta=\max\{\sup(C_{\tau}\cap\beta_{k_1})\mid\tau\in\tr(\beta_{k_1},\beta)\}.$$
By $\alpha_0'=\alpha_0$ and $\sup(C_{\alpha'_j}\cap\alpha^*)\le \eta<\eta+1\le\alpha^*$ for all $j<k_2$,
we get that $\alpha_{j+1}'=\min(C_{\alpha_{j'}}\bks\alpha^*)=\min(C_{\alpha_j}\bks\eta+1)=\alpha_{j+1}$ for all $j<k_2$.
In particular, $m\ge k_2$ and $\alpha_{k_2}=\alpha^*$.
It now follows from $\psi(\Delta(\alpha,\beta))=(k_1,k_2)$ and
the definition of $\rts$, that $\rts(\alpha,\beta)=(\alpha^*,\beta^*)$.
\end{proof}
\end{proof}
\subsection{The countable cofinality case}
In this subsection, we prove Theorem \ref{t12}, for the case that $\lambda$ is a singular cardinal of countable cofinality.
Here, instead of appealing to Theorem \ref{31}, we shall make use of Eiswroth's theorem concerning \emph{off-center club guessing}.
\begin{thm}[Eisworth, \cite{eisworth4}] Suppose that $\lambda>\cf(\lambda)=\omega$.
Then there exists an increasing sequence of successor cardinals
$\overrightarrow \mu=\langle \mu_m\mid m<\omega\rangle$, and a matrix
$\overrightarrow e=\langle e_\delta^m \mid \delta\in E^{\lambda^+}_{\omega_1}, m<\omega\rangle$ such that:
\begin{enumerate}
\item $\{ e^m_\delta\mid m<\omega\}$ is an increasing chain of club subsets of $\delta$;
\item $|e^m_\delta|\le\mu_m$ for all $\delta$ and $m$;
\item for every club $E\s\lambda^+$, there exist stationarily many $\delta\in E^{\lambda^+}_{\omega_1}$
such that $\sup(e^m_\delta\cap E\cap E^{\delta}_{>\mu_m})=\delta$ for all $m<\omega$;
\item $\sup\overrightarrow\mu=\lambda$.
\end{enumerate}
\end{thm}
The next lemma is analogous to that of Lemma \ref{32}.
\begin{lemma}[Eisworth, \cite{eisworth4}] Let $\lambda$ and $\overrightarrow e,\overrightarrow\mu$ be as in the preceding theorem.
Then there exists a matrix $\overrightarrow C=\langle C^m_\alpha\mid \alpha<\lambda^+, m<\omega\rangle$ such that for every
$\alpha<\lambda^+$:
\begin{itemize}
\item $\max(C^m_{\alpha+1})=\alpha$ for all $m<\omega$;
\item $\{C^m_\alpha\mid m<\omega\}$ is a chain of club subsets of $\alpha$;
\item $|C^m_\alpha|\le\max\{\mu_m,\cf(\alpha)\}$;
\item if $\delta\in C^m_\alpha\cap E^{\lambda^+}_{\omega_1}$, then $e^m_\delta\s C^m_\alpha$.
\end{itemize}
\end{lemma}
From now on, we fix a cardinal $\lambda$ and sequences
$\overrightarrow e,\overrightarrow C, \overrightarrow \mu$ as above.
\begin{notation} if $\sigma\in{}^{<\omega}\omega$ is a non-empty sequence and $m<\omega$,
we define $\ell(\sigma):=\dom(\sigma)$, $\sigma^\frown m$ as $\sigma\cup\{(\ell(\sigma),m)\}$,
and $\lift(\sigma)$ as the unique element of ${}^\omega\omega$ satisfying for all $n<\omega$,
$$\lift(\sigma)(n)=\begin{cases}\sigma(n),&n\in\dom(\sigma)\\
\sigma(\max(\dom(\sigma)),&\text{otherwise}\end{cases}.$$
\end{notation}
In this subsection, we shall be interested in conducting \emph{generalized walks}:
for $\alpha<\beta<\lambda^+$ and $\sigma\in{}^{<\omega}\omega$, let $\beta_0:=\beta$, and $\beta_{n+1}:=\min(C^{\lift(\sigma)(n)}_{\beta_n}\bks\alpha)$
whenever $n<\omega$ and $\beta_n>\alpha$.
The outcome $\beta=\beta_0>\ldots>\beta_{k+1}=\alpha$ is considered as \emph{the
$(\overrightarrow C,\sigma)$-walk from $\beta$ down to $\alpha$}, and we denote:
\begin{itemize}
\item $\rho_\sigma(\alpha,\beta):=k+1$;
\item $\trs(\alpha,\beta):=\{ (\lift(\sigma)(i),\beta_i)\mid i\le k\}$;
\item $\Tr_\sigma(\alpha,\beta):=\{ (\lift(\sigma)(i),\beta_i)\mid i\le k+1\}$.
\end{itemize}
Fix a surjection $\psi:\omega\rightarrow{}^{<\omega}\omega\times{}^{<\omega}\omega$ such that the pre-image of every element is infinite.
Let $(\overrightarrow\lambda,\overrightarrow f)$ be
a scale for $\lambda$, and let, as in the previous subsection, for all $\alpha<\beta<\lambda^+$:
$$\Delta(\alpha,\beta):=\sup\{ i<\cf(\lambda)\mid f_\alpha(i)\ge f_\beta(i)\}.$$
For $\alpha<\beta<\lambda^+$, we now define $\rts(\alpha,\beta)$. Let $\sigma_1,\sigma_2$ be such that
$\psi(\Delta(\alpha,\beta))=(\sigma_1,\sigma_2)$.
Next, let:
\begin{itemize}
\item $\beta_0>\dots>\beta_{n}$ denote the
$(\overrightarrow{C},\sigma_1)$-walk from $\beta$ down to $\alpha$;
\item $\eta:=\max\{\sup(C^i_{\tau}\cap\beta_{\ell(\sigma_1)})\mid (i,\tau)\in\Tr^\circ_{\sigma_1}(\beta_{\ell(\sigma_1)},\beta)\}$;
\item $\alpha_0>\ldots>\alpha_{m}$ denote the
$(\overrightarrow{C},\sigma_2)$-walk from $\alpha$ down to $\eta+1$;
\end{itemize}
We would like to assign $\rts(\alpha,\beta):=(\alpha_{\ell(\sigma_2)},\beta_{\ell(\sigma_1)})$.
Of course, if $\ell(\sigma_1)> n$, $\alpha<\eta+1$, or $\ell(\sigma_2)> m$,
then the above definition would not make sense. In such cases, we just put $\rts(\alpha,\beta):=(\alpha,\beta)$.
The next definition is parallel to Definition \ref{dddd}.
\begin{defn} For $A\s\lambda^+$, $\eta\le\epsilon<\lambda^+$, and $\sigma\in{}^{<\omega}\omega$, let
$$A_{\sigma,\eta}(\epsilon):=\left\{\alpha\in A\mid \rho_\sigma(\epsilon,\alpha)=\ell(\sigma), \max\{\sup(C^i_\tau\cap\epsilon)\mid (i,\tau)\in\trs(\epsilon,\alpha)\}=\eta\right\}.$$
\end{defn}
\begin{lemma}\label{34} If $A,B$ are cofinal subsets of $\lambda^+$, and $\theta<\lambda$,
then the following set is stationary:
$$S^{A,B}_{>\theta}:=\{\epsilon\in E^{\lambda^+}_{>\theta}\mid
\exists(\sigma_1,\sigma_2,\eta)\in{}^{<\omega}\omega{}\times^{<\omega}\omega\times\epsilon\left( |B_{\sigma_1,\eta}(\epsilon)|=|A_{\sigma_2,\eta}(\epsilon)|=\lambda^+\right)\}.$$
\end{lemma}
\begin{proof} Suppose that $A,B$ are cofinal subsets of $\lambda^+$, and $\theta<\lambda$.
Let $E$ be an arbitrary club subset of $\lambda^+$,
and let us prove that $S^{A,B}_{>\theta}\cap E\not=\emptyset$.
Let $S$ denote the set of all $\delta\in E^{\lambda^+}_{\omega_1}$ for which there exists
an elementary submodel $M\prec\mathcal H(\chi)$
such that:
\begin{itemize}
\item $\{\overrightarrow C,A,B\}\s M$;
\item $M\cap\lambda^+=\delta>\lambda$;
\item $\sup(e^m_\delta\cap E\cap E^{\delta}_{>\mu_m})=\delta$ for all $m<\omega$.
\end{itemize}
Then $S$ is a stationary set, and $D:=\{\varepsilon<\lambda^+\mid \sup(S\cap\varepsilon)=\varepsilon\}$ is a club.
By the choice of the club guessing sequence $\overrightarrow e$,
let us fix some $\nu\in E^{\lambda^+}_{\omega_1}$ such that $\sup(e^m_\nu\cap E\cap E^\nu_{>\mu_m})=\nu$ for all $m<\omega$.
Let $\beta:=\min(B\bks\nu+1)$, and $\alpha:=\min(A\bks\nu+1)$.
For all $m<\omega$, let
\begin{itemize}
\item $\beta=\beta^m_0>\ldots>\beta^m_{n(m)+1}=\nu$ denote the $(\overrightarrow{C},\langle m\rangle)$-walk
from $\beta$ down to $\nu$;
\item $\alpha=\alpha^m_0\ldots>\alpha^m_{k(m)+1}=\nu$ denote the $(\overrightarrow{C},\langle m\rangle)$-walk
from $\alpha$ down to $\nu$.
\end{itemize}
Clearly, the sequence $\langle \beta_0^m\mid m<\omega\rangle$ is a
constant sequence. As $\{C^m_\beta\mid m<\omega\}$ is an increasing chain,
the sequence $\langle \min(C^m_\beta\bks\nu)\mid m<\omega\rangle$
stabilizes, and hence the sequence $\langle \beta^m_1\mid
m<\omega\rangle$ is eventually constant.
Of course, an iterative application of this last observation yields that the sequence
$\langle \rng(\trm(\nu,\beta)) \mid m<\omega\rangle$ is eventually constant.
By the same kind of consideration, we get that the sequence
$\langle \rng(\trm(\nu,\alpha)) \mid m<\omega\rangle$ is eventually constant.
Thus, let us pick some large enough $m^*$ such that $\langle (\rng(\trm(\nu,\beta)),\rng(\trm(\nu,\alpha)))\mid m^*\le m<\omega\rangle$ is a
constant sequence.
Next, pick an even larger $m<\omega$ so that $m>m^*$, and
$$\mu_m\ge \max\{\cf(\beta^{m^*}_{n(m^*)}),\cf(\alpha^{m^*}_{k(m^*)})\}.$$
For notational simplicity, denote $n:=n(m), k:=k(m), \beta_i:=\beta^m_i, \alpha_j:=\alpha^m_j$.
Put
$$\gamma:=\max(\{\sup(C^m_{\beta_i}\cap\nu)\mid i<n\}\cup \{\sup(C^m_{\alpha_j}\cap\nu)\mid j<k\}).$$
Since $\gamma<\nu$, let us pick a large enough $\varepsilon\in e^m_\nu\cap D$
such that $\varepsilon>\gamma$ and $\cf(\varepsilon)>\mu_m$.
\begin{claim}\label{291} \begin{enumerate}
\item $\trm(\varepsilon,\beta)=\{(m,\beta_i)\mid i\le n\}$;
\item $\trm(\varepsilon,\alpha)=\{(m,\alpha_j)\mid j\le k\}$.
\end{enumerate}
\end{claim}
\begin{proof}
As $(\gamma,\nu)\cap C^m_{\beta_i}=\emptyset$ for all $i<n$, and
$\varepsilon\in(\gamma,\nu)$, we get that
$\min(C^m_{\beta_i}\bks\varepsilon)=\min(C^m_{\beta_i}\bks\nu)$ for all
$i<n$, and hence $\rng(\Tr_{\langle m\rangle}(\varepsilon,\beta))\bks\beta_n=\{\beta_i\mid i\le n\}$. Likewise, by $\varepsilon\in(\gamma,\nu)$
and $(\gamma,\nu)\cap C^m_{\alpha_j}=\emptyset$ for all $j<k$, we
get that $\rng(\Tr_{\langle m\rangle}(\varepsilon,\alpha))\bks\alpha_k=\{\alpha_j\mid j\le k\}$.
Since $\nu\in C^m_{\beta_n}\cap C^m_{\alpha_k}\cap E^{\lambda^+}_{\omega_1}$, we
have $e^m_\nu\s C^m_{\beta_n}\cap C^m_{\alpha_k}$, which implies
$\varepsilon\in C^m_{\beta_n}\cap C^m_{\alpha_k}$.
So $\rho_{\langle m\rangle}(\varepsilon,\beta)=n+1$,
$\rho_{\langle m\rangle}(\varepsilon,\alpha)=k+1$, and the assertion of the claim follows immediately.
\end{proof}
Put $$\gamma':=\max\{\sup(C^i_{\tau}\cap\varepsilon)\mid
(i,\tau)\in\trm(\varepsilon,\beta)\cup\trm(\varepsilon,\alpha)\}.$$
Since $\gamma<\varepsilon<\nu$, we have $\gamma'=\max\{\gamma,\sup(C^m_{\beta_n}\cap\varepsilon),\sup(C^m_{\alpha_k}\cap\varepsilon)\}$.
Recalling the choice of $m$, and the fact that $|C^i_\tau|\le\max\{\mu_i,\cf(\tau)\}$ for
all $i$ and $\tau$, we infer that:
$$\max\{|C^{m}_{\beta_{n}}|,|C^{m}_{\alpha_{k}}|\}\le\max\{\mu_{m},\cf(\beta^{m^*}_{n(m^*)}),\cf(\alpha^{m^*}_{k(m^*)})\}=\mu_m.$$
By the choice of $\varepsilon$, we have $\mu_m<\cf(\varepsilon)$,
and hence $\sup(C^m_{\beta_n}\cap\varepsilon)<\varepsilon$,
$\sup(C^m_{\alpha_k}\cap\varepsilon)<\varepsilon$. Altogether, $\gamma'<\varepsilon$.
Consider the constant functions $\sigma_1:n+1\rightarrow\{m\}, \sigma_2:k+1\rightarrow\{m\}$,
and note that $\ell(\sigma_1)=\rho_{\langle m\rangle}(\varepsilon,\beta)$, $\ell(\sigma_2)=\rho_{\langle m\rangle}(\varepsilon,\alpha)$.
Since $\varepsilon\in e^m_\nu\cap D$, we now pick a large enough $\delta\in S$ satisfying $\gamma'<\delta<\varepsilon$,
and consider additional walks. For all $w<\omega$, let:
\begin{itemize}
\item $\beta=\beta^{\cdot w}_0>\ldots>\beta^{\cdot w}_{n(w)+1}=\delta$ denote the $(\overrightarrow{C},\sigma_1{}^\frown w)$-walk
from $\beta$ down to $\delta$;
\item $\alpha=\alpha^{\cdot w}_0>\ldots\alpha^{\cdot w}_{k(w)+1}=\delta$ denote the $(\overrightarrow{C},\sigma_2{}^\frown w)$-walk
from $\alpha$ down to $\delta$.
\end{itemize}
\begin{claim}\label{292} For every $w<\omega$:\begin{enumerate}
\item $\Tr^\circ_{\sigma_1{}^\frown w}(\delta,\beta)=\trw(\delta,\varepsilon)\cup\trm(\varepsilon,\beta)$;
\item $\Tr^\circ_{\sigma_2{}^\frown w}(\delta,\alpha)=\trw(\delta,\varepsilon)\cup\trm(\varepsilon,\alpha)$.
\end{enumerate}
\end{claim}
\begin{proof} Fix $w<\omega$.
By definition of $\gamma'$ and since $\delta\in(\gamma',\varepsilon)$, we get that $\beta^{\cdot w}_i=\beta_i$ for all $i\le n$,
and hence $\beta^{\cdot w}_{n+1}=\varepsilon$. So $\rng(\trm(\varepsilon,\beta))=\{\beta^{\cdot w}_0,\ldots\beta^{\cdot w}_n\}$,
and $\rng(\trw(\delta,\varepsilon))=\{\beta^{\cdot w}_{n+1},\ldots,\beta^{\cdot w}_{n(w)}\}$.
Likewise, $\alpha^{\cdot w}_j=\alpha_j$ for all $j\le k$, and hence $\alpha^w_{k+1}=\varepsilon$.
So $\rng(\trm(\varepsilon,\alpha))=\{\alpha^{\cdot w}_0,\ldots,\alpha^{\cdot w}_k\}$,
and $\rng(\trw(\delta,\varepsilon))=\{\alpha^{\cdot w}_{k+1},\ldots,\alpha^{\cdot w}_{k(w)}\}$.
\end{proof}
Fix a large enough $w^*<\omega$ for which
$\langle \rng(\trw(\delta,\varepsilon)))\mid w^*\le w<\omega\rangle$ is a
constant sequence.
Also, let:
\begin{itemize}
\item $\zeta:=\min(\rng(\Tr^\circ_{\langle w^*\rangle}(\delta,\varepsilon)))$;
\item $\gamma_\alpha:=\max\{\sup(C^i_{\tau}\cap\delta)\mid (i,\tau)\in \Tr^\circ_{\sigma_2{}^\frown w^*}(\delta,\alpha), \delta\not\in C^i_\tau\}$;
\item $\gamma_\beta:=\max\{\sup(C^i_{\tau}\cap\delta)\mid (i,\tau)\in \Tr^\circ_{\sigma_1{}^\frown w^*}(\delta,\beta), \delta\not\in C^i_\tau\}$.
\end{itemize}
Evidently, $\gamma_\alpha<\delta$, $\gamma_\beta<\delta$.
Now, pick a large enough $w<\omega$ so that $w>w^*$, and $\mu_w>\cf(\zeta)$.
Then, pick a large enough $\epsilon\in e^w_\delta\cap E$ such that
$\sup(e^w_\delta\cap\epsilon)>\max\{\gamma_\alpha,\gamma_\beta\}$, and
$\cf(\epsilon)>\max\{\mu_w,\theta\}$.
\begin{claim}\label{293} $\Tr^\circ_{\sigma_1{}^\frown w}(\epsilon,\beta)=\Tr^\circ_{\sigma_1{}^\frown w}(\delta,\beta)$, and
$\Tr^\circ_{\sigma_2{}^\frown w}(\epsilon,\alpha)=\Tr^\circ_{\sigma_2{}^\frown w}(\delta,\alpha)$.
\end{claim}
\begin{proof} As in the proof of Claim \ref{291}.
\end{proof}
Denote:
\begin{itemize}
\item $\eta_\alpha:=\max\{\sup(C^i_\tau\cap\epsilon)\mid (i,\tau)\in\Tr^\circ_{\sigma_2{}^\frown w}(\epsilon,\alpha)\}$;
\item $\eta_\beta:=\max\{\sup(C^i_\tau\cap\epsilon)\mid (i,\tau)\in\Tr^\circ_{\sigma_1{}^\frown w}(\epsilon,\beta)\}$.
\end{itemize}
By definition of $\gamma_\beta,\eta_\beta,\zeta,w,w^*$, Claims \ref{292}, \ref{293},
and the fact that $\gamma_\beta<\epsilon<\delta$,
we get that $\eta_\beta=\max\{\gamma_\beta,\sup(C^w_\zeta\cap\epsilon)\}$.
Since $\delta\in C^w_\zeta$, we get that $\sup(C_\zeta^w\cap\epsilon)\ge\sup(e^w_\delta\cap\epsilon)>\gamma_\beta$,
and hence $\eta_\beta=\sup(C^w_\zeta\cap\epsilon)$.
Likewise, $\eta_\alpha=\sup(C^w_\zeta\cap\epsilon)$.
Denote $\eta:=\sup(C^w_\zeta\cap\epsilon)$. Then we have just established that
\begin{itemize}
\item $\beta\in B_{\sigma_1',\eta}(\epsilon)$, for $\sigma_1':=\lift(\sigma_1{}^\frown w)\restriction \rho_{\sigma_1{}^\frown w}(\epsilon,\beta)$, and
\item $\alpha\in A_{\sigma_2',\eta}(\epsilon)$, for $\sigma_2':=\lift(\sigma_2{}^\frown w)\restriction \rho_{\sigma_2{}^\frown w}(\epsilon,\alpha)$.
\end{itemize}
By the choice of $\delta$, let us now pick an elementary submodel $M\prec\mathcal H(\chi)$
such that $\{A,B,\overrightarrow C\}\s M$ and $M\cap\lambda^+=\delta$.
The sets $B_{\sigma_1',\eta}(\epsilon)$ and $A_{\sigma_2',\eta}(\epsilon)$ are definable from parameters in $M$,
and hence belongs to $M$. Since $|M|=\lambda$ and $\beta\in B_{\sigma_1',\eta}(\epsilon)\bks M$,
we get that $|B_{\sigma'_1,\eta}(\epsilon)|=\lambda^+$.
Likewise, $|A_{\sigma_2',\eta}(\epsilon)|=\lambda^+$.
By the choice of $\epsilon$,
we have $\cf(\epsilon)>\mu_w>\cf(\zeta)$, and hence $\cf(\epsilon)>\max\{\mu_w,\cf(\zeta)\}\ge|C^w_\zeta|$.
So $\eta<\epsilon$, and hence
$\epsilon\in S^{A,B}_{>\theta}$.
Finally, recalling that $\epsilon\in e^w_\delta\cap E$, we conclude that
$S^{A,B}_{>\theta}\cap E\not=\emptyset$.
\end{proof}
\begin{thm} For every cofinal subsets $A,B$ of $\lambda^+$, and every $\theta<\lambda$,
there exists a stationary subset $S^*\s E^{\lambda^+}_{>\theta}$ such that $\rts[A\circledast B]\supseteq [S^*]^2$.
\end{thm}
\begin{proof}[Proof sketch] This is very much like the proof of Theorem \ref{t25}.
Namely, given $A,B$ and $\theta$, we appeal to Lemma \ref{34} and Fodor's lemma,
and find $\sigma_1,\sigma_2\in{}^{<\omega}\omega$ and $\eta<\lambda^+$ such that the following set is stationary:
$$S:=\{\epsilon\in E^{\lambda^+}_{>\theta}\mid |B_{\sigma_1,\eta}(\epsilon)|=|A_{\sigma_2,\eta}(\epsilon)|=\lambda^+\}.$$
We then pick a continuous $\in$-chain of relevant elementary submodels $\langle M_i\mid i<\lambda^+\rangle$,
let $S^*:=\{ \epsilon\in S\mid M_\epsilon\cap\lambda^+=\epsilon\}$,
and argue that $\rts[A\circledast B]\supseteq [S^*]^2$.
\end{proof}
\section{Conclusions}
Building on the above and Eisworth's \cite{eisworth2},\cite{eisworth1}, we get the next theorem:
\begin{thm}\label{t31} For every singular cardinal, there exists a function $\rtts:[\lambda^+]^2\rightarrow[\lambda^+]^2\times\cf(\lambda)$
such that if $\langle a_\alpha\mid \alpha<\lambda^+\rangle$ is a family of pairwise disjoint members of $[\lambda^+]^{<\cf(\lambda)}$,
then there exists a stationary set $S\s\lambda^+$ such that for all $\langle \alpha,\beta,\delta\rangle\in [S]^2\times \cf(\lambda)$,
there exists $\alpha<\beta<\lambda^+$ such that $D[t_\alpha\times t_\beta]=\{\langle\alpha,\beta,\delta\rangle\}$.
\end{thm}
\begin{proof} In \cite{eisworth1}, Eisworth proved that there exists a function $D:[\lambda^+]\rightarrow[\lambda^+]^2\times\cf(\lambda)$
such that for every cofinal $A\s \lambda^+$, there exists a stationary subset $S\s\lambda^+$
such that $D``[A]^2\supseteq [S]^2\times\cf(\lambda)$.
Shortly afterwards, in a subsequent paper, Eisworth proved \cite{eisworth2}
that there exists a function $D':[\lambda^+]\rightarrow\lambda^+\times\lambda^+\times\cf(\lambda)$
such that if $\langle a_\alpha\mid \alpha<\lambda^+\rangle$ is a family of pairwise disjoint members of $[\lambda^+]^{<\cf(\lambda)}$,
then there are stationary subsets $S$ and $T$ of $\lambda^+$, such that for all $\langle \alpha,\beta,\delta\rangle\in S\circledast T\times \cf(\lambda)$,
there exists $\alpha<\beta<\lambda^+$ such that $D'[a_\alpha\times a_\beta]=\{\langle\alpha,\beta,\delta\rangle\}$.
Thus, given $\alpha<\beta<\lambda^+$, let
$\rtts(\alpha,\beta):=D(\rts(\alpha',\beta'))$, where $D'(\alpha,\beta)=(\alpha',\beta',\delta')$.
It is easy to see that $\rtts$ has the required properties.
\end{proof}
\begin{remark}
The assertion of Theorem \ref{t31} may be strengthened to require that $S$ contains ordinals of arbitrarily high cofinality.
\end{remark}
We are now in a position to prove the Introduction's Theorem \ref{t11}.
\begin{thm}\label{t32} For every singular cardinal $\lambda$, and every $\theta\le\lambda^+$, the following are equivalent:
\begin{enumerate}
\item $\lambda^+\not\rightarrow[\lambda^+]^2_\theta$;
\item $\lambda^+\not\rightarrow[\lambda^+;\lambda^+]^2_\theta$;
\item $\pr_1(\lambda^+,\lambda^+,\theta,\cf(\lambda))$.
\end{enumerate}
\end{thm}
\begin{proof}
$(1)\Rightarrow (2)$. Suppose that $f:[\lambda^+]^2\rightarrow\theta$ is a function witnessing $\lambda\not\rightarrow[\lambda^+]^2_\theta$.
Define a function $g:[\lambda^+]^2\rightarrow\theta$ by stipulating that $g(\alpha,\beta):=f(\rts(\alpha,\beta))$. Then $g$ witnesses
$\lambda\not\rightarrow[\lambda^+;\lambda^+]^2_\theta$.
$(2)\Rightarrow (3)$. Suppose that $f:[\lambda^+]^2\rightarrow\theta$ is a function witnessing $\lambda\not\rightarrow[\lambda^+;\lambda^+]^2_\theta$.
Define a function $g:[\lambda^+]^2\rightarrow\theta$ by stipulating that $g(\alpha,\beta):=f(\alpha',\beta')$,
where $\rtts(\alpha,\beta)=(\alpha',\beta',\delta)$, and $\rtts$ is the function given by the preceding theorem.
Then $g$ witnesses $\pr_1(\lambda^+,\lambda^+,\theta,\cf(\lambda))$.
$(3)\Rightarrow (1)$ Suppose that $f:[\lambda^+]^2\rightarrow\theta$ is a function witnessing $\pr_1(\lambda^+,\lambda^+,\theta,\cf(\lambda))$.
Then $f$ witnesses $\lambda\not\rightarrow[\lambda^+]^2_\theta$, as well.
\end{proof}
As an immediate corollary, we get an affirmative answer to the question raised by Eisworth in \cite{eisworth3}.
\begin{cor} If $\lambda$ is a singular cardinal and $\pr_1(\lambda^+,\lambda^+,\lambda^+,\cf(\lambda))$ fails,
then any collection of less than $\cf(\lambda)$ many stationary subsets of $\lambda^+$ must reflect simultaneously.
\end{cor}
\begin{proof} By \cite{eisworth3}, if $\lambda^+\not\rightarrow[\lambda^+]^2_{\lambda^+}$ fails,
then any collection of less than $\cf(\lambda)$ many stationary subsets of $\lambda^+$ must reflect simultaneously.
Now, appeal to Theorem \ref{t32}.
\end{proof}
\bibliographystyle{plain}
|
\section{Introduction}
Observing that a physical system is only known up to some finite precision, Andronov and Pontriaguyn \cite{Andronov} suggested in 1937 that the study of dynamical systems should focus on stable systems, i.e., those that do not change under small perturbations. Rather strikingly, it has turned out that the topologically $C^1$-stable dynamics (also called structurally stable systems) can be analyzed;
indeed, they are exactly the uniformly hyperbolic systems~\cite{Mane88} that satisfy an additional technical assumption. However, the structurally stable diffeomorphisms are not dense
in the $C^1$-topology and therefore the study of such systems is insufficient even up to an arbitrarily small perturbation.
Hence, much of the focus of current research in dynamical systems is to extend
our understanding beyond uniform hyperbolicity, in the hope of eventually obtaining
a global theory of ``most" systems (see for example \cite{Palis}).
A natural approach is to consider weaker forms of stability. First, one should
design a stability property that holds for interesting examples outside of uniform hyperbolicity.
Second, one should establish large sets of stable dynamics by finding robust
global phenomena enforcing this property. Third, if this stability fails to hold densely,
one should look for robust local mechanisms responsible for its failure.
\footnote{This dichotomy of phenomena and mechanisms has been put forward by Pujals
\cite{Pujals} and is related to Palis conjectures for a global picture of dynamics \cite{Palis}.}
The obvious candidate for such a notion is $C^r$-stability with $r>1$, but it turns out to be very difficult to study because of a lack of perturbation lemmas. We keep the $C^1$-topology but replace topological conjugacy by a looser, entropy-based notion and realize the first step of the above program.
\subsection{Stability of the large entropy measures for Bonatti-Viana diffeomorphisms}
We will show that many constructions, including the Bonatti-Viana diffeomorphisms which we will describe below, have strong stability properties, though they are not hyperbolic and therefore not structurally stable.
We first review some standard definitions. Let $\Diff^1(M)$ denote the space of $C^1$-diffeomorphisms of a compact manifold $M$ endowed with some Riemannian structure. $\Diff^1(M)$ is endowed with its usual topology and distance $d_{C^1}(f,g)$
(see \cite[Section 8.1.1]{RobinsonBook}).
A map $f:M\rightarrow M$ is {\it transitive} if there exists some $x\in M$ whose forward orbit is dense in $M$. A diffeomorphism $f:M\rightarrow M$ is {\it $C^1$-robustly transitive} if there exists a neighborhood $\mathcal{U}$ of $f$ in $\mathrm{Diff}^1(M)$ such that each $g\in\mathcal{U}$ is transitive.
A {\it Borel isomorphism} of $\psi:X\to Y$ is a bijection such that $\psi$ and $\psi^{-1}$ are both Borel maps. Let $\Prob(f)$ be the set of invariant Borel probability
measures of $f$.
The topological entropy of a system $(X,f)$, denoted $h_{\mathrm{top}}(f)$, is a number that measures the topological complexity of the system. On the other hand, if $\mu\in\Prob(f)$, then the measure theoretic entropy, denoted $h_{\mu}(f)$, of a dynamical system is a number that measures the complexity of the system as seen by the measure $\mu$. (See~\cite[Sections 3.1 and 4.3]{KH} for precise definitions.)
The variational principle states that if $f$ is a continuous self-map of a compact metrizable space, then $h_{\mathrm{top}}(f)=\sup_{\mu\in\Prob(f)}h_{\mu}(f)$, see for instance~\cite[p. 181]{KH}. A measure $\mu\in\Prob(f)$ such that $h_{\mathrm{top}}(f)=h_{\mu}(f)$ is a {\it measure of maximal entropy}. If there is a unique measure of maximal entropy, then $f$ is called {\it intrinsically ergodic}.
\begin{defn}\label{def:h-stable}
A diffeomorphism $f:M\to M$ is \new{entropically $C^r$-stable} if
for every $g\in \Diff^r(M)$ that is $C^r$-close to $f$ there exists
a Borel isomorphism $\psi:M'\to M''$ where $M', M''$ are Borel subsets of $M$ such that the following properties hold:
\begin{itemize}
\item $\psi\circ g = f\circ \psi$ on $M'$,
\item $\tilde h(f,M\setminus M'') := \sup \{ h(f,\nu) :
\mu\in\Prob(f|_{M\setminus M''})\}< h_\top(f)$, and
\item $\tilde h(g,M\setminus M') < h_\top(g)$.
\end{itemize}
\end{defn}
It is convenient to call \emph{large entropy measures,} the \emph{ergodic} invariant probability measures with entropy greater than some constant $h$ (strictly less than the topological entropy). We denote their set by $\Prob_\erg^h(f)$. Hence, a system is entropically $C^r$-stable if its large entropy measures ``stay the same" for any sufficiently close diffeomorphism.
We shall establish that large entropy measures also maintain the following property, enjoyed by the Anosov system (for which we can take $X_1=X$):
\begin{defn}
Let $f:X\to X$ be a Borel isomorphism of a metric space $X$ and $\mathcal M$ be a set of invariant probability measures of $f$.
We say that $\mathcal M$ is {\bf almost expansive} if there exists a number $\eps_0>0$ such that, for every $\mu\in\mathcal M$, for $\mu$-a.e. $x\in M$, for all $y\in M$ $\sup_{n\in\ZZ} d(f^nx,f^ny)\geq \eps_0$ implies $x=y$.
\end{defn}
What can we say about the low entropy measures? On the one hand, a recent result of Hochman \cite{Hochman} shows that entropy-conjugacy with an Anosov system implies Borel conjugacy to this Anosov system up to subsets of zero measure with respect to any aperiodic invariant probability measure. This gives the corresponding strengthening of the stability property. On the other hand, expansivity can fail dramatically around low entropy measures. The theory of symbolic extension and entropy structures of Boyle and Downarowicz~\cite{BD, Dow} allows to precisely define such phenomena.
Recall that a symbolic system is defined as the left shift $(x_n)_{n\in\ZZ}\mapsto
(x_{n+1})_{n\in\ZZ}$ acting on a shift invariant, compact subset of $\mathbb N^{\mathbb Z}$
(not necessarily of finite type).
Given a homeomorphism $f$ of a compact metrizable space $X$ a
{\it symbolic extension} of $(X,f)$ is a continuous surjection
$\varphi\colon \Sigma\to X$ such that $f\circ\varphi=\varphi\circ \sigma$
and $(\Sigma,\sigma)$ is a symbolic system. Boyle and Downarowicz~\cite{BD} have shown that for a given dynamical system there is a natural connection between how entropy arises on finer and finer scales "around" an invariant probability measure and the existence of such symbolic extensions that approximate, closely in entropy, the system equipped with that measure. We shall see that, for some examples of entropically stable systems considered in this paper, the dynamics at low entropy does not only fail to be expansive but prevents the existence of any symbolic extension.
We will study diffeomorphisms $f:M\to M$ of a compact manifold $M$ with a weak form of hyperbolicity called a dominated splitting.
A $Df$-invariant splitting of the tangent bundle of some invariant set $\Lambda$
$$
T_\Lambda M=E_1\oplus\cdots\oplus E_k
$$
is \emph{dominated} if each bundle has constant dimension (at least two of them non-zero) and there exists an integer $\ell\geq1$ with the following property. For every $x\in\Lambda$, all $i=1,\dots,(k-1)$, and every pair of unitary vectors $u\in E_1(x)\oplus\dots\oplus E_i$ and $v\in E_{i+1}(x)\oplus\dots\oplus E_k(x)$, it holds that
$$
\frac{|Df_x^\ell (u)|}{|Df^{\ell}_x (v)|}\leq \frac{1}{2}.
$$
(See for example~\cite[Appendix B, Section 1]{BUH} for properties of systems with a dominated splitting.) A diffeomorphism $f\in\mathrm{Diff}^1(M)$ is {\it partially hyperbolic} if there exists a dominated splitting $TM=E^s \oplus E^c\oplus E^u$ where $E^s$ is uniformly contracting, $E^u$ is uniformly expanding (at least one of $E^s$ and $E^u$ is non-trivial). The diffeomorphism $f$ is {\it strongly partially hyperbolic} if $E^s$ and $E^u$ are both non-trivial. A strongly partially hyperbolic diffeomorphism is {\it hyperbolic} (or Anosov) if $TM=E^s\oplus E^u$. The (stable) index is the dimension $\dim E^s$.
The first goal of the present paper is to analyze a class diffeomorphisms described by Bonatti and Viana~\cite{BV}. This was the first example of a robustly transitive diffeomorphism that is not partially hyperbolic.
\begin{thm}\label{thm:applyToBV}
There exist a $C^\infty$-diffeomorphism, $f$, of the $4$-torus, $\TT^4$, and an open set $\mathcal{U}\subset \mathrm{Diff}^1(M)$ containing $f$ such that each $g\in \mathcal{U}$ is $C^1$-robustly transitive, not partially hyperbolic and are all entropy-conjugate to the same Anosov diffeomorphism.
Moreover, for each $g\in\mathcal U$, the large entropy measures are almost expansive. Nevertheless, there exist a non-empty open set $\mathcal{V}\subset \mathcal{U}$ and a $C^1$-residual set $\mathcal{D}\subset \mathcal{V}$ such that each $g\in \mathcal{D}$ has no symbolic extension.
\end{thm}
\begin{cor}
All diffeomorphisms $g\in\mathcal U$ above are not partially hyperbolic, are not structurally stable or $\Omega$-stable but are $C^1$-entropically stable.
\end{cor}
In particular, this gives a nonempty open set of nonpartially hyperbolic diffeomorphisms
with constant topological entropy and unique measures of maximal entropy that define pairwise isomorphic measure-preserving transformations. We derive this from an abstract result (Theorem \ref{thm:main}).
\begin{rem}
The diffeomorphisms $g\in\mathcal D$ defined above are far from entropy-expansive (i.e., they do not satisfy $h_\loc(f,\eps)=0$ for any $\eps>0$). Indeed, asymptotic entropy-expansivity, i.e., $\lim_{\eps\to0} h_\loc(g,\eps)=0$ would imply the existence of a symbolic extension with nice properties (see Sec. \ref{sec:non-concentration} for definitions). Nevertheless, these diffeomorphisms satisfy:
$$h_\top(g)=h_\top(g,\eps_0)$$
for $\eps_0>0$ the implicit constant in the definition of almost expansivity (which can even be chosen independently of $g\in\mathcal V$).
\end{rem}
\subsection{Previous Results}\label{sec:previous}
Newhouse and Young \cite{NY83} proved that a class of partially hyperbolic, nonhyperbolic robustly transitive diffeomorphisms\footnote{This class was described originally by Shub~\cite{shub}.} that are $C^0$ deformations of Anosov diffeomorphisms are $C^1$-entropically stable. More precisely, they showed that these diffeomorphisms have a unique measure of maximal entropy that is isomorphic to the measure of maximal entropy for the Anosov system but this can be strengthened to the entropic stability defined above.
Together with M. Sambarino and V. Vasquez the authors of the present work proved the following related result:
\begin{thm}~\cite{bfsv}~\label{t.bsfv} For any $d\geq 3$, there exists a nonempty open set $\mathcal{U}$
in $\mathrm{Diff}(\mathbb T^d)$ satisfying:
\begin{itemize}
\item each $f\in \mathcal{U}$ is strongly partially hyperbolic, robustly transitive, and
$C^1$-entropically stable;
in particular the topological entropy is locally constant at $f$;
\item each $f\in \mathcal{U}$ has equidistributed periodic points; and
\item no $f\in \mathcal{U}$ is Anosov or structurally stable.
\end{itemize}
\end{thm}
We note that the abstract result below (Theorem \ref{thm:main}) gives another proof of the first point of the above theorem.
\medbreak
Using different techniques, F.~Rodriguez Hertz, J.~Rodriguez Hertz, Tahzibi, and Ures~\cite{RHRHTU} have obtained a rather precise description of the following systems:
\begin{thm}\cite{RHRHTU}
Consider an accessible partially hyperbolic diffeomorphism of a 3-dimensional manifold having compact center leaves. Either it has a unique entropy maximizing measure with zero center Lyapunov exponent,
or it has a positive, even number of entropy maximizing ergodic, invariant measures, all of them with nonzero center Lyapunov exponent.
\end{thm}
We refer to their paper for definitions of the terms above. Notice that this shows that transitivity, entropic stability and non-uniqueness of the entropy maximizing measures can robustly coexist. Also, it shows that uniqueness of the entropy maximizing measure does not hold generically outside of the uniformly hyperbolic systems, even assuming topological transitivity.
\subsection{Further Questions}
There are a number of natural questions that follow from Theorem~\ref{thm:applyToBV}.
\subsubsection{A phenomenon for entropic stability}
After this analysis of a large class of examples, we would like to find the phenomena responsible for entropic stability and in particular formulate more general conditions, instead of specifying a perturbative scheme. The following condition might be sufficient:
\begin{defn} \cite{BuzziICMP}
Let $f\in\Diff(M)$. Define
$$
h^{k}(f):=\sup\{h_\top(f,\phi([0,1]^{k})): \phi\in C^\infty(\RR^k,M)\}.
$$
We say that a dominated splitting $TM=E^{cs}\oplus E^{cu}$ is \new{entropy-hyperbolic} if the following holds:
\begin{itemize}
\item $h^{\dim E^{cu}-1}(f)< h_\top(f)$, and
\item $h^{\dim E^{cs}-1}(f^{-1}) < h_\top(f)$.
\end{itemize}
\end{defn}
\begin{que}
Does the entropy-hyperbolicity of a dominated splitting imply the finiteness of the number of entropy maximizing ergodic invariant measures? Does it implies entropic-stability?
\end{que}
We think that the answer to the first question above is affirmative, on the basis of partial results assuming stronger versions of the condition.
\begin{rem}
Entropy-hyperbolicity, as formulated above, seems to be excessively restrictive. For instance, the disjoint union of two Anosov systems with distinct indices cannot satisfy it for trivial reasons. Also, we do not know of a system with an entropy-hyperbolic dominated splitting which is not given by an isotopy from an Anosov system.
\end{rem}
\subsubsection{Entropy-conjugacy to uniform systems}
The stability observed by Newhouse and Young and in this work, follows from a weak form of conjugacy to a uniformly hyperbolic system: almost conjugacy (topological conjugacy up to negligible sets for the maximal entropy measure of each system) or entropy-conjugacy (see below Def. \ref{def:h-conjugate}). One might think that such a conjugacy is actually the rule rather than the exception. They formulated this idea as follows:
\begin{que}\cite{NY83}\label{q.ny1} For any compact manifold $M$ and $r\geq1$, let $\mathcal{B}(M)$ denote the set of $C^r$ diffeomorphisms such that
\begin{enumerate}
\item $f$ has finitely many ergodic, invariant measures of maximal entropy, and
\item on the support of each such measure, $f$ is almost conjugate to some Axiom A diffeomorphism.
\end{enumerate}
Is $\mathcal{B}(M)$ residual in $\mathrm{Diff}^r(M)$?
\end{que}
One can of course reformulate this question, replacing the above notion of almost conjugacy by entropy-conjugacy (almost conjugacy does not imply and is not implied by entropy-conjugacy).
\subsubsection{Generic stability and finiteness}
\begin{que}\label{q.ny} Does a generic diffeomorphism admit finitely many entropy-maximizing ergodic and invariant measures? Is it entropically stable?
\end{que}
However, for all we know, entropic stability (like structural stability) could fail to be dense. More precisely,
S. Crovisier~\cite{Cro} suggested that the diffeomorphisms with homoclinic
classes robustly without a dominated splitting might provide a (large) set of points of
variation of the topological entropy and thus give a negative answer to Question \ref{q.ny}.
\section{Abstract Result}\label{s.abstract}
In this section, we state Theorem \ref{thm:main}, from which Theorem~\ref{thm:applyToBV} will be deduced in section \ref{s.BV}. Theorem \ref{thm:main} is our main result. It proves that a large class of deformations of Anosov
diffeomorphisms that are big in the $C^1$ topology, do not modify large entropy measures. We first state the somewhat technical assumptions as three definitions and then the theorem. We conclude this section with an outline of the proof.
For simplicity, we assume that $M=\mathbb T^d$, the $d$-dimensional torus and leave the obvious modifications necessary to deal with the (slightly) more general manifolds carrying Anosov systems to the interested reader. Also, Anosov or Anosov system will mean Anosov diffeomorphism.
\medbreak
The first requirement will ensure that the perturbation is $C^1$-small, except possibly on a union of a given number of well-separated balls of small radius:
\begin{defn}\label{def:sparse}
An \new{$(\eps,N)$-sparse deformation} (or just: $(\eps,N)$-deformation) of $f\in\Diff^1(M)$ is a diffeomorphism $g:M\rightarrow M$
such that there exist $x_1,\dots,x_N\in M$
and $r>0$ satisfying:
\begin{itemize}
\item $d_{C^1}(g|_{M\setminus B_r},f|_{M\setminus B_r})<\eps$ where $B_r:=\bigcup_{i=1}^N B(x_i,r)$;
\item $r<\eps$;
\item $\min_{i\ne j} d(x_i,x_j)>\eps^{1/2}$.
\end{itemize}
$B_r$ is called a \new{strong support} of the deformation and $r$ is called its \new{radius}.
\end{defn}
Before stating the second requirement, we need to recall some facts about cone conditions and hyperbolicity. The {\it cone} $C^1_\alpha$ of aperture $\alpha>0$ defined by a decomposition $E^1\oplus E^2$ of a Euclidean space $E$ is:
$$
C^1_\alpha:=\{ v^1+v^2\in E: v^i\in E^i\text{ and }\|v^2\|\leq\alpha\|v^1\| \}.
$$
$\dim(E^1)$ is called the dimension of the cone. For a manifold $M$, a {\it cone field} is the specification of a cone $C(x)$ of fixed dimension in each $T_xM$, $x\in M$. A boundaryless submanifold $\Sigma$ is {\it tangent to a cone field} $C$ if
\begin{enumerate}
\item[(i)] the submanifold and the cone have the same (constant)
dimension and
\item[(ii)] $T_x\Sigma\subset C(x)$ at every $x\in M$.
\end{enumerate}
Let $f$ be Anosov with an adapted Riemannian metric: there exists a $Df$-invariant continuous splitting $TM=E^s\oplus E^u$ with the following bounds:
$$\begin{array}{rlll}
\lambda_0& =\min_{x\in M}\min_{v\in E^u_x\setminus\{0\}}\frac{\| Df v\|}{\| v\|}>1 \textrm{ and }\\
\mu_0& =\max_{x\in M}\max_{v\in E^s_x\setminus\{0\}}\frac{\| Df v\|}{\| v\|}<1.
\end{array}$$
($\lambda_0$ is the minimum expansion for $Df$ in the unstable direction, and $\mu_0$ is the minimum contraction for $Df$ in the stable direction). The hyperbolicity strength is:
$$
\lambda_1=\min\{\lambda_0, \mu_0^{-1}\}>1.
$$
Let $C^u_\alpha$ and $C^s_\alpha$ denote the cones defined by the hyperbolic splitting of $f$ associated to $E^s\oplus E^u$ as above, for an aperture $\alpha>0$ to be determined.
\begin{rem} Observe that the cones above are those defined by $f$, not by $\tilde f$. These will be the only cone fields that we consider.
\end{rem}
We now formulate our second requirement on the deformations. It keeps the dominated splitting, even inside the strong support.
\begin{defn}A diffeomorphism $g:M\rightarrow M$
\new{($\alpha, \rho, \Lambda$)-respects the domination of $f$} if it satisfies the following for all $x\in M$ and all $y,z\in B(x,\rho)$ such that:
$$
y-x\in C^u_{\alpha}(x) \text{ and }g(z)-g(x)\in C^s_{\alpha}(g(x))
$$
then
\begin{enumerate}
\item $\|g(y)-g(x)\|/\|y-x\|> \Lambda \|g(z)-g(x)\|/\|x-z\|$
\item $g(y)-g(x)\in C^u_{\alpha}(g(x))$ and $z-x\in C^s_{\alpha}(x)$
\end{enumerate}
\end{defn}
This assumption of non-linear domination will ensure that large center-unstable disks are mapped by $g$ to similar disks and will be used to build invariant center-unstable foliations (and likewise for center-stable ones). The point of the above definition is to make the scale $\rho>0$ explicit.
Our third (and last) requirement is that even if a vector in the center-unstable direction can be contracted, this contraction is weak (and analogously in the center-stable direction):
\begin{defn}
For $\gamma>0$ a diffeomorphism $g:M\rightarrow M$ is \new{$\gamma$-nearly hyperbolic} with respect to a dominated splitting
$TM=E^{cu}\oplus E^{cs}$ if for some $C\in(1, \infty)$ and for all $n\geq0$ the following conditions are satisfied:
\begin{enumerate}
\item[(i)] $\|Dg^n v^{cu}\|\geq C^{-1}e^{-\gamma n}$ for all $ v^{cu}\in E^{cu}$ and
\item[(ii)]
$ \|Dg^n v^{cs}\|\leq C e^{\gamma n}$ for all $v^{cs}\in E^{cs}$.
\end{enumerate}
\end{defn}
We will be interested in systems that are $\gamma$-nearly hyperbolic for $\gamma$ near zero.
\medbreak
The following notion introduced in \cite{BuzziSIM} describes the type of conjugacy we will obtain. Essentially the dynamics are conjugate with respect to ergodic invariant measures with large entropy:
\begin{defn}\label{def:h-conjugate}
Two dynamical systems $f:X\to X$ and $g:Y\to Y$ are \new{entropy-conjugate}
if there exists a partially defined bimeasurable bijection: $\psi:Y\setminus Y_0
\to X\setminus X_0$ such that:
\begin{itemize}
\item $\psi\circ g = f\circ \psi$ on $Y\setminus Y_0$,
\item $\tilde h(f,X_0) := \sup \{ h(f,\nu) :
\mu\in\Prob(f,X_0)\}< h_{\mathrm{top}}(f)$, and
\item $\tilde h(g,Y_0) < h_{\mathrm{top}}(g)$.
\end{itemize}
\end{defn}
We shall prove that the systems we consider are entropy conjugate to Anosov systems.
\begin{rem}\label{r.conjugate}
Entropic stability of $f$ means that
any diffeomorphism $C^1$-close to $f$ is entropy-conjugate to $f$.
\end{rem}
\medbreak
We now can state our main result. Theorem~\ref{thm:applyToBV} will follow directly from this result.
\begin{thm}\label{thm:main}
Let $f:M\to M$ be an Anosov diffeomorphism on $M=\TT^d$, $d\geq2$ and let $N\geq1$ be some integer. There exists $t:=t(f,N)>0$ with the following property.
Let $\eps,\alpha,\gamma \in(0,t)$ and let $\rho:= (\epsilon^{1/2}-2\epsilon)\cdot\mathrm{diam}M$ and $\Lambda>\frac{\epsilon^{1/2} + 2\epsilon}{\epsilon^{1/2}-2\epsilon}>1$.
Any $g\in\Diff^1(M)$ satisfying:
\begin{enumerate}
\item[(H1)] $g$ is an $(\eps,N)$-sparse deformation of $f$;
\item[(H2)] $g$ ($\alpha, \rho,\Lambda$)-respects the domination of $f$;
\item[(H3)] $g$ is $\gamma$-hyperbolic;
\end{enumerate}
is entropy-conjugate to $f$. Moreover, the set of large entropy measures, $\Prob_\erg^h(g)$ for some $h<h_\top(g)$, is almost expansive.
\end{thm}
\subsection{Strategy of proof}
The proof of Theorem~\ref{thm:main} splits
into the following steps:
\begin{enumerate}
\item Existence of canonical, invariant, center-stable and
center-unstable foliations for $g$. The existence of these will follow from the dominated splitting
and the respect of the domination of $f$ at a certain scale (Sec.~\ref{sec:foliations}).
\item Under the factor map on the Anosov dynamics defined by the shadowing property, the measure-theoretic entropy can decrease only slightly (Sec.~\ref{sec:extensionAnosov}).
\item The large entropy measures of $g$ give little mass to the strong support of the deformation (Sec.~\ref{sec:non-concentration}).
\item The factor map is actually an entropy-conjugacy, proving Theorem \ref{thm:main} (Sec.~\ref{sec:factor}).
\end{enumerate}
\section{Invariant foliations}\label{sec:foliations}
The goal of this section is to build center-stable and center-unstable invariant foliations for our deformation $g$ of some Anosov system $f\in\Diff^1(M)$. As above, we restrict ourselves for simplicity to the case where $M=\TT^d$.
We first recall some definitions.
\medbreak
A \emph{continuous foliation} $\FF$ of dimension $k$ with $C^r$ leaves is a partition of the manifold such that there is locally an homeomorphism mapping $\FF$ to the partition of $\mathbb R^d$ into $k$-planes for some $0\leq k\leq d$ and such that its restriction to any such plane is $C^r$.\footnote{It is well-known that even in the hyperbolic case, the local homeomorphisms mapping the stable (or unstable) leaves to planes cannot always be chosen $C^1$.}
It is well-known that, in full generality, the existence of a dominated splitting does \emph{not} imply the existence of invariant foliations tangent to each sub-bundle. In fact even with the stronger assumption of partial hyperbolicity the center direction may not be integrable. In~\cite{BW, P} there are discussions on the integrability of the bundles and some classical examples are given where the integrability does not hold.
Let $D$ be a smooth open disk embedded in $M$. Its \emph{inner radius} at some point $x\in D$, is the distance between $x$ and $\partial D:=\overline{D}\setminus D$.
\begin{thm}\label{thm:foliations}
Let $f\in\Diff^1(M)$ be Anosov with hyperbolic strength $\lambda>1$. Let $1<\Lambda<\lambda$ and $\alpha>0$. There exists $\eps_1(f,\Lambda,\alpha)>0$ such that for all $0<\eps<\eps_1$ the following holds.
Let $N\geq1$. and set $\rho=(\eps^{1/2}-2\epsilon)$.
Let $g\in\Diff^1(M)$ be an $(\eps,N)$-deformation of $f$ which $(\alpha, \rho,\Lambda)$-respects the domination of $f$.
Then $g$ has a dominated splitting $TM=E^{cs}\oplus E^{cu}$ with the same index as the hyperbolic splitting of $f$. Moreover, $g$ admits a center-stable foliation, $\mathcal F^{cs}$, and a center-unstable foliation, $\mathcal F^{cu}$, with the following properties:
\begin{enumerate}
\item each foliation is continuous with $C^1$-leaves;
\item the leaves of the foliations are everywhere tangent to $E^{cs}$, $E^{cu}$ respectively;
\item the foliations are invariant under $g$.
\end{enumerate}
\end{thm}
The following non-shrinking property is key to our construction:
\begin{cla}\label{lem:domination}
Let $f,g\in\Diff^1(M)$ and $N\geq 1$, $\Lambda>1,\alpha>0$, $\eps>0$, and $\rho:=(\eps^{1/2}-2\epsilon)$. Assume that $g$ is an $(\eps,N)$-deformation of an Anosov $f$ which $(\alpha,\rho,\Lambda)$-respects the domination of $f$. Assume also $\Lambda> \frac{\epsilon^{1/2} +2\epsilon}{\epsilon^{1/2}-2\epsilon}>1$.
Let $x\in M$. Then for any disk $D^u_{\rho}$ tangent to $C^u_{\alpha}$ and with inner radius $\rho$ at $x$, $g(D^u_{\rho})$ contains a disk tangent to $C^u_{\alpha}$ and with inner radius at least $\rho$ at $g(x)$.
\end{cla}
\begin{proof}[Proof of Claim \ref{lem:domination}]
Let $x\in M$. The invariance of the center-unstable cone implies that $g(D^u_\rho(x))$ is tangent to $C^u_\alpha$. It remains to see that this disk also has large diameter. As $f$ and $g$ are $C^0$-close,
$d(f(y),g(y))\leq 2\epsilon\cdot\mathrm{diam}M.$
It follows that
$$
\begin{array}{llll}
d(g(x),\partial g( D^u_\rho(x))) &\geq d(f(x),\partial f(D^u_\rho(x)))-4\epsilon\\
& \geq \Lambda\rho-4\epsilon\\
& >\frac{\epsilon^{1/2}+2\epsilon}{\epsilon^{1/2}- 2\epsilon}(\epsilon^{1/2}-2\epsilon)-4\epsilon\\
&=\rho.
\end{array}$$
The claim is proved.
\end{proof}
\medskip
\begin{proof}[Proof of Theorem \ref{thm:foliations}]
We observe that the existence of the dominated splitting $E^{cs}\oplus E^{cu}$ for $g$ is a well-known consequence of the existence of the invariant cones. See for instance~\cite[p. 293]{BUH}.
We first fix $x\in M$ and construct a sequence of disks of inner radius at least $\rho$ at $x$ and everywhere tangent to \modified{$C^{cu}$.
For each $n\geq0$, let $D_{x,n}^{-n}$ be an embedded open
smooth disk tangent to the unstable cone field $C^u_\alpha$ defined by the hyperbolic splitting of $f$ and with inner radius $\rho$ at $g^{-n}(x)$. By compactness of $M$, there is $\eps_0>0$, independent of $x$ so that, for any $0<\eps<\eps_0$, such a disk always locally exists, independently of any integrability condition.\footnote{For instance, we can take a disk in the local stable manifold of $x$ with respect to the Anosov diffeomorphism, $f$.} By reducing $\eps_0>0$ if necessary, we also ensure $\Lambda> \frac{\epsilon_0^{1/2} +2\epsilon_0}{\epsilon_0^{1/2}-2\epsilon_0} >1$. Using an obvious identification, we can chose $D_{x,n}^{-n}$ to be the graph of a map defined on an open subset of $E^u_\alpha(x)$ taking values in $E^{cs}_\alpha(x)$ and with small Lipschitz constant.}
Let $D_{x,n}^{k+1}:=g(D_{x,n}^k)\cap B(g^{k+1}x,\rho)$ for all
$k=-n,\dots,-1$. Standard graph transform estimates show that
$(D_{x,n}^0)_{n\geq0}$ is a family of graphs of equicontinuous functions.
Moreover, their domains of definition contain the disk of
$x+E^{cu}_x$ with center $x$ and radius $\rho$ from Claim~\ref{lem:domination}. Thus, one can find a subsequence such that these functions converge uniformly to a function with bounded Lipschitz constant. Let $D_x$ be the limit graph.
Note that every $y\in D_x$ satisfies $d(g^{-n}y,g^{-n}x)\leq \rho$ for all $n\geq0$. This allows the use the non-linear domination and to get, through standard arguments, that $D_x$ is $C^1$
with tangent spaces obtained as intersections of nested and
exponentially shrinking cones. In particular, these tangent
spaces coincide with $E^{cu}$.
Let us show that the sequence $(D_{x,n}^0)_{n\geq0}$ is actually
convergent by checking that the limit graph is unique
and independent of the choice of $D_{x,n}^{-n}$:
\begin{figure}[htb]
\begin{center}
\psfrag{x}{$x$}
\psfrag{y}{$y$}
\psfrag{z}{$y'$}
\psfrag{d}{$D_1$}
\psfrag{e}{$D_2$}
\psfrag{f}{$\Delta$}
\psfrag{g}{$C^{s}$}
\psfrag{h}{$C^{u}$}
\includegraphics{linear.eps}
\caption{unique disks}\label{f.linear}
\end{center}
\end{figure}
By contradiction, consider two distinct limit disks $D_1$ and $D_2$.
Thus, there exists a disk, $\Delta$, tangent to $E^{s}$ which intersects
$D_1$ and $D_2$ in two distinct points $y$ and $y'$. By construction, for all $n\geq0$ we have
\begin{itemize}
\item $g^{-n}y, g^{-n}y'\in B(g^{-n}x,\rho)$;
\item $g^{-n}y-g^{-n}x\in C^{cu}_\alpha(g^{-n}x)$;
\item $g^{-n}y-g^{-n}y'\in C^{cs}_\alpha(g^{-n}x)$.
\end{itemize}
It follows that, for all $n\geq0$:
$$
\frac{\|y-x\|}{\|g^{-n}y-g^{-n}x\|} \geq \Lambda^{n} \frac{\|y-y'\|}{\|g^{-n}y-g^{-n}y'\|}.
$$
But $g^{-n}D,g^{-n}D'$ are contained in the cone $C^{u}_{\alpha_1}(g^{-n}x)$, thus
$$\|g^{-n}y-g^{-n}y'\|\leq K\|g^{-n}y-g^{-n}x\|$$ for some uniform $1<K<\infty$
and
$$
\|y-y'\| \leq \Lambda^{-n} \frac{\|g^{-n}y-g^{-n}y'\|}{\|g^{-n}y-g^{-n}x\|} \|y-x\|
\leq K\Lambda^{-n} \|y-x\|.
$$
Letting $n\to\infty$, we see that $y=y'$, a contradiction.
Note that the canonical character of the disks $D_x$ imply their equivariance: $g(D_x)\cap B(gx,\rho)=D_{gx}$. Also, the above argument implies the following uniqueness property.
For any $x,y\in M$,
if $z\in D^{cu}_x\cap D^{cu}_y$, then
\begin{equation}\label{eq:Dcu-uniq}
D^{cu}_x\cap D^{cu}_y\cap B(z,\rho)\subset D^{cu}_z.
\end{equation}
We now define the partition candidate to be an invariant center-unstable foliation. For each $x\in M$ we let $\FF^{cu}(x)$ be the set of all $y\in M$ such that
there exist finitely many points $x_1,\dots,x_n$ satisfying: $x\in D^{cu}_{x_1}$, $y\in D^{cu}_{x_n}$ and $D^{cu}_{x_i}\cap D^{cu}_{x_{i+1}}\ne\emptyset$ for $i=1,\dots,n-1$. It follows from this definition that $\FF^{cu}$ is a partition and that it is invariant: $g(\FF^{cu}(x))=\FF^{cu}(g(x))$.
To prove that $\FF^{cu}$ is indeed a foliation, it remains to check that each $\FF^{cu}(x)$ intersects any small ball in a disjoint union of smooth disks and that the connected component of $x$ depends continuously in the $C^1$ topology of the base point $x$.
Let us set $F_x:=\FF^{cu}(x)\cap B(x,\rho/2)$. Obviously,
$$F_x=\bigcup_{y\in F_x} D^{cu}_y\cap B(x,\rho/2).$$
It follows from \eqref{eq:Dcu-uniq} that this is a disjoint union in the sense that either
$$D^{cu}_y\cap B(x,\rho/2)=D^{cu}_{y'}\cap B(x,\rho/2)$$
or the two sets are disjoint. Thus, the connected component of $F_x$ containing $x$ is $D^{cu}_x\cap B(x,\rho/2)$. The construction of $D^{cu}_x$ shows that this is indeed a $C^1$ submanifold that depends continuously on $x$.
The claims of the theorem for $\FF^{cu}$ are proved. The proofs for $\FF^{cs}$ are completely analogous.
\end{proof}
\section{Almost Principal Extension of the Anosov}\label{sec:extensionAnosov}
In this section we let $g\in\mathrm{Diff}^1(M)$ ($M=\TT^d$), a sufficiently small $C^0$-perturbation of an Anosov diffeomorphism, $f$, and study the continuous factor map $\pi:(M,g)\to (M,f)$ given by the shadowing lemma (see Lemma \ref{l.shadowing} below).
We observe that the fibers $\pi^{-1}(x)$ for $x\in M$ have a small diameter. Second, we show that if $g$ respects the domination of $f$ and is nearly hyperbolic, then for a.e. $x\in M$, $\pi^{-1}(x)$ is contained in a leaf of the center-unstable or center-stable foliation (given by Theorem~\ref{thm:foliations}).
\subsection{Shadowing for Anosov diffeomorphisms}
We recall the following well-known fact about hyperbolic dynamics. For a proof see for instance~\cite[p. 109]{shub}.
\begin{lem}[Shadowing Lemma]\label{l.shadowing}
Let $f:M\to M$ be Anosov. There exist numbers $\eps_0>0$ and $K_0<\infty$ with the following property.
For any homeomorphism $g:M\to M$ with
$$d_{C^0}(f,g):=\sup_{x\in M} d(f(x),g(x))+d(f^{-1}(x),g^{-1}(x))<\eps_0,$$
there is a topological factor map $\pi:(M,g)\to(M,f)$. Moreover, $\sup_{x\in M} \diam(\pi^{-1}(x))\leq K_0 d_{C^0}(f,g)$.
\end{lem}
To fix some notations, we recall the following classical notion.
\begin{defn}
Two foliations $\FF^{1},\FF^{2}$ have a \new{(local) product structure} if there exist constants $\tau_1,\tau_2>0$ and $1<K<\infty$ such that the following hold:
for all points $x,y$ within distance less than $\tau_1$, $\FF^{1}_{\tau_2}(x)$
(the connected component of $\FF^{1}(x)\cap B(x,\tau_2)$ containing
$x$) intersects the similarly defined $\FF^{2}_{\tau_2}(x)$ at
exactly one point, $z$, and $d(x,z)\leq K d(x,y)$.
\end{defn}
\begin{rem}A compact manifold with transverse continuous foliations $\mathcal{F}_1$ and $\mathcal{F}_2$ with $C^1$ leaves has a product structure for $\mathcal{F}_1$ and $\mathcal{F}_2$ for \emph{some} constants $\tau_1,\tau_2,K$.
\end{rem}
\subsection{Inclusion in $\FF^{cu}$ or $\FF^{cs}$}
The next proposition shows that for an appropriate deformation $g$, the fibers of the ergodic invariant probability measures for $g$ disintegrated over $f$ are contained in the leaves of one of the dynamical foliations.
\begin{prop}\label{prop:flat}
Let $f:M\to M$ be Anosov with shadowing constants $\eps_0>0$ and $K_0<\infty$ and hyperbolicity strength $\lambda>1$. Let $N\geq1$ and $1<\Lambda<\lambda$, $\alpha>0$, $\tau_1,\tau_2>0$ and $K<\infty$. There exists $\eps_2(f,\Lambda,\alpha,\eps_0,K_0,\tau_1,\tau_2,K)>0$ with the following property for all $0<\eps<\eps_2$ and $g\in\Diff^1(M)$ which
\begin{itemize}
\item is an ($\eps,N$)-sparse deformation of $f$;
\item ($\alpha,\rho,\Lambda)$-respects the domination with $\rho:=(\eps^{1/2}-2\eps)$;
\item preserves center-stable and center-unstable foliations $\FF^{cs},\FF^{cu}$ tangent to the cone fields $C^s_\alpha,C^u_\alpha$ and define a product structure with constants $\tau_1,\tau_2,K$.
\end{itemize}
For any any $g$-invariant, ergodic probability measure $\nu$, there exists $\sigma=cs$ or $cu$ such that,
$$
\text{ for $\nu$-a.e. $x\in M$ } \nu_x(\FF^\sigma_{\tau_2}(x)) = 1,
$$
where $\nu=\int_M \nu_x \, d\pi_*\nu$ is the Rokhlin disintegration of $\nu$
w.r.t. $\pi$ (see \cite{Rudolph}).
\end{prop}
\begin{proof}
We shall establish the required property under a (finite) number of upperbounds on $\eps_2$. Recall the number $\eps_1(f,\Lambda,\alpha)>0$ from Theorem \ref{thm:foliations}. The first bound is:
$$
\eps_2<\min\{\eps_1, 1/({2K_0})^{2}\}
$$
so $2\epsilon K_0<\epsilon^{1/2}$: the balls of radius $K_0\epsilon$ around the $N$ centers of the $(\epsilon, N)$-deformation $g$ are disjoint. This will be useful with regards to the Shadowing Lemma: recall that $\diam(\pi^{-1}(x))\leq K_0\eps$.
Let $\nu$ be an invariant ergodic measure for $g$ with its disintegration $(\nu_x)_{x\in M}$ as above. Since $\pi$ is a semi-conjugacy, $\pi_*\nu$ is an ergodic probability measure for $f$.
It is convenient to set aside the trivial case where $\nu_x=\delta_x$ for a
set of positive (and hence full) $\pi_*\nu$-measure of points $x\in M$.
Let $\mu$ be the Cartesian square of $\nu$ relatively to the $\pi$ factor.
In other words, $\mu$ is the probability measure for $g\times g$ on $M\times M$
given by
$$
\mu = \int_M \nu_x\times\nu_x \, d\pi_*\nu.
$$
Observe that it is $g\times g$-invariant.
We define a measurable function $R:M\times M\to [0,\infty]$ as follows.
For $(x,y)\in M\times M$, let
$z$ be the unique intersection point
of $\FF^{cs}_{\tau_2}(x)$ and $\FF^{cu}_{\tau_2}(y)$ (if one exists). We set
$$
R(x,y) := \left\{
\begin{array}{llll}
\frac{d_{\FF^{cs}(x)}(x,z)}{d_{\FF^{cu}(y)}(y,z)} &\textrm{ if }z\textrm{ exists and }z\neq y\\
\infty & \textrm{ else}
\end{array}
\right.
$$
where $d_N(\cdot,\cdot)$ denotes the geodesic distance along the
submanifold $N$ using the induced Riemannian structure.
Note that as $\pi(x)=\pi(y)$ for $\mu$-a.e.~$(x,y)$,
$d(x,y)<K_0\epsilon$. To use the product structure, we impose our second bound on $\eps_2$:
$$
\eps_2<\tau_1/K_0
$$
Thus, by the transversality assumption on $\FF^{cu},\FF^{cs}$, $z$ is well-defined and $x,y\in B(z,KK_0\epsilon)$ .
To use the respect of the domination, we impose our third bound:
$$
\eps_2< (2/KK_0)^2,
$$
so $\rho>KK_0\epsilon$ and therefore, if $R(x,y)<\infty$, then $R(g^nx,g^ny)\to0$ when $n\to\infty$.
The invariance of $\mu$ implies $R(x,y)=0$ or $\infty$ $\mu$-a.e.
We claim that $R$ is $\pi$-measurable. Otherwise there would
exist a set of positive $\pi_*\nu$-measure of points $x\in M$,
such that
$$\begin{array}{llll}
\{(y,z)\in M\times M:R(y,z)=0\}\textrm{ and}\\
\{(y,z)\in M\times M:R(y,z)=\infty\}
\end{array}
$$
have both positive $\nu_x\times\nu_x$-measure.
Now, observe that
$$
\{(y,z):R(y,z)=0\} = \{(y,z): \FF^{u}_{\tau_2}(y)=\FF^{u}_{\tau_2}(z)\}
$$
and that if this set has positive $\nu_x\times\nu_x$-measure for a set of $x\in M$ with positive $\pi_*\nu$-measure, then there exists a measurable function of $x$, $y_x$ such that $\nu_x(\FF^{cu}_{\tau_2}(y_x))>0$ over a set of positive $\pi_*\nu$-measure. Similarly there exists a measurable $z_x$ such that $\nu_x(\FF^{su}_{\tau_2}(z_x))>0$. It follows that:
$$
(\nu_x\times\nu_x)(\FF^{cu}_{\tau_2}(y_x)\times\FF^{cs}_{\tau_2}(z_x))>0.
$$
As $\nu_x\ne\delta_x$ by assumption, it follows that $0<R(y,z)<\infty$
with positive $\mu$-measure, a contradiction.
\end{proof}
\section{Non-concentration}\label{sec:non-concentration}
We consider an asymptotically entropy-expansive diffeomorphism (whose definition is recalled below) and show that its large entropy measures cannot be concentrated around a fixed number of points. We will apply this to Anosov diffeomorphisms.
We recall Bowen's entropy formula for a subset $Y\subset M$ in terms of dynamical $(\eps,n)$-balls $$B_f(x,\eps,n):=\{y\in M:\forall 0\leq k<n\; d(f^ky,f^kx)<\eps\}.$$ We have
$
h_\top(f,Y) := \lim_{\eps\to0} h_\top(f,Y,\eps)$
($h_\top(f)=h_\top(f,M)$) with
$$
h_\top(f,Y,\eps):=\limsup_{n\to\infty}\frac1n\log r_f(\eps,n,Y)
$$
where $r_f(\eps,n,Y)$ is the minimal number of dynamical $(\eps,n)$-balls needed to cover $Y$.
Katok \cite{Katok} established a similar formula for the entropy of an ergodic invariant probability measure:
$$
h(f,\mu) = \lim_{\eps\to0} h_\top(f,\mu,\eps) \text{ with } h_\top(f,\mu,\eps):=\limsup_{n\to\infty}\frac1n\log r_f(\eps,n,\mu)
$$
where $r_f(\eps,n,\mu)$ is the minimal number of dynamical $(\eps,n)$-balls with union of measure at least $1/2$. \footnote{One can replace $1/2$ by any other fixed number in $(0,1)$.}
Finally, we recall Misiurewicz's local (or conditional, or tail) entropy \cite{MisiurewiczLocal}:
$$
h_\loc(f) := \lim_{\eps\to0} h_\loc(f,\eps) \text{ with }
h_\loc(f,\eps):= \sup_{x\in M} h_\top(f,B_f(x,\eps,\infty)).
$$
\subsection{Large entropy measures of $f$}
\begin{lem}\label{lem:deconcentrate-f}
Let $f$ be a homeomorphism of $M$ which is asymptotically $h$-expansive (i.e., $h_\loc(f)=0$) with $h_\top(f)>0$.
For any $\eta>0$ and $N\geq1$, there exist $h<h_\top(f)$ and $r>0$ such that any $\mu\in\Prob_\erg^h(f)$ satisfies $\mu\left(\bigcup_{i=1}^N B(x_i,r)\right)<\eta$, for any set of $N$ points $x_1,\dots,x_N\in M$.
\end{lem}
\begin{proof}
Let $0<\eta<1$. Pick $0<\eps<\eta h_\top(f)/4$. As $f$ is asymptotically $h$-expansive we know
there exists a constant $s_0>0$ such that, for any $s>0$, any subset $Y$, and any $n\geq0$ the following holds
$$
r(s,n,Y) \leq C(s,s_0) e^{\eps n} r(s_0,n,Y).
$$
Also $r(s_0/3,n,M)\leq C_0 e^{(h_\top(f)+\eps)n}$ for some $C_0<\infty$ and all
$n\geq0$.
Observe that, for any $s>0$, any integer $n\geq0$ and a decomposition $n=n_1+\dots+n_k$ into a sum of positive integers
we have
$$
r(3s,n,Y) \leq \prod_{i=1}^{k} r(s,n_i,f^{n_1+\dots+n_{i-1}}Y).
$$
To see this, consider the map $\iota:x\mapsto (y_1,\dots,y_{k})$ where
$$
f^{n_1+\dots+n_{i-1}}(x)\in B_f(y_i,s,n_i)
$$
with the $y_i$'s taken from a minimal set of centers of $(s,n)$-balls making a cover of $f^{n_1+\dots+n_{i-1}}Y$. Take a minimal set $C_1$ such that the $\{B_f(x,s,n)\}_{x\in C_1}$ is a cover of $Y$. Select a minimal subset $C_2\subset C_1$ such that $\iota:C_2\to\iota(C_1)$ is a bijection. Clearly the cardinality of $C_2$ satisfies the above bound. We claim that $\{B_f(x,3s,n)\}_{x\in C_2}$ is a cover of $Y$. This follows from the fact that $\iota(x')=\iota(x)$ implies $B_f(x',s,n)\subset B_f(x,3s,n)$.
Now let $n_0<\infty$ satisfy
$$\frac{\log N+\log C_0}{n_0}<\eps\textrm{ and }\binom{2[n/n_0]+2}{n}\leq e^{\eps n/2}$$
for all $n\geq 0$. Let $r>0$ be such that $B(x,r)\subset B_f(x,r_0/2,n_0)$.
Fix $N$ points $x_1,\dots,x_N$ and $B_r:=B(x_1,r)\cup\dots\cup B(x_N,r)$. Let $\mu$ be an invariant ergodic measure of $f$ with $\mu(B_r)>\eta$.
We now bound the entropy of $\mu$ by estimating the number of $(s_0,n)$-balls necessary to cover some set $M'$ of measure more than $1/2$.
Observe that for a typical $x$ and $n$ large enough, we can decompose the integer interval
$[0,n[$ into subintervals, half of them being of the form $[a,a+n_0[$ with $f^ax\in
B(x_i,r)$ and the sum of their lengths at least $\eta n$. Therefore, we have
$$
r(s_0,n,M') \leq \sum_{n_1+\dots+n_k+kn_0=n} \prod_{i=1}^k r(s_0/3,n_i,M) r(s_0/3,n_0,B_r)
$$
The previous estimates and the Birkhoff ergodic theorem yield a subset $M'$ of $M$ with $\mu(M')>1/2$ such that,
for all large $n$:
$$\begin{array}{llll}
\frac1n\log r(s_0,n,M') \leq &\eta \log N/n_0 + (1-\eta) (h_\top(f)+\eps) + \\
& \log C_0/n_0
+ \log \binom{2[n/n_0]+1}{n}/n.
\end{array}
$$
It follows that
$$\begin{aligned}
h(f,\mu) &\leq h(f,\mu,s_0)+\eps \\
& \leq (\eta \log N+\log C_0)/n_0 + (1-\eta) (h_\top(f)+\eps) +2\eps \\
& \leq h:=h_\top(f)+3\eps -\eta h_\top(f)
< h_\top(f).
\end{aligned}$$
\end{proof}
\section{Proof of Theorem~\ref{thm:main}}\label{sec:factor}
Let $N\geq1$ be an integer and let $f$ be an Anosov diffeomorphism of a compact manifold $M$.
\subsection{Choice of the numbers $\alpha,\eps>0$ and $\Lambda>1$}
We endow $M$ with an adapted Riemannian metric. Let $\eps_*,K_*$ be the two numbers as in the shadowing Lemma (Lemma \ref{l.shadowing}). Let $\lambda>1$ be the hyperbolicity strength of $f$.
We fix $\Lambda\in (1,\lambda)$ and pick $\alpha>0$ and $0<\eps_0<\eps_*/2$ small enough so that $\Lambda>\frac{\epsilon_0^{1/2} +2\epsilon_0}{\epsilon_0^{1/2}-2\epsilon_0}>1$ and, for all $\tilde f\in\Diff^1(M)$ with $d_{C^1}(\tilde f,f)<\eps_0$, for all $x\in M$:
$$\begin{aligned}
&\forall v\in C^u_{\alpha}(x)\qquad \| D\tilde fv\|\geq\Lambda\|v\| \text{ and }D\tilde f v\in C^u_\alpha(\tilde fx) \\
&\forall v\in C^s_{\alpha}(x)\qquad \| D\tilde fv\|\leq\Lambda^{-1}\|v\| \text{ and }D\tilde f^{-1} v\in C^s_\alpha(\tilde f^{-1}x)
\end{aligned}$$
where $C^u_\alpha,C^s_\alpha$ are the cone fields with aperture $\alpha$ around the unstable and stable bundles of $f$. We also fix $R_0>0$ such that, for all $x\in M$, for all $y\in (x+C^u_{\alpha}(x))\cap B(x,R_0)$
\begin{equation}\label{eq:exp-Anosov1}
\| \tilde fy-\tilde fx\|\geq\Lambda\|y-x\|
\text{ and }\tilde f y-\tilde f x\in C^u_\alpha(\tilde f x)\\
\end{equation}
and, likewise, if $y\in (x+C^s_{\alpha}(x))\cap B(x,R_0)$
\begin{equation}\label{eq:exp-Anosov2}
\| \tilde f^{-1}y-\tilde f^{-1}x\|\geq\Lambda\|y-x\|
\text{ and }\tilde f^{-1}y-\tilde f^{-1} x\in C^s_\alpha(\tilde f^{-1}x).
\end{equation}
Observe that by compactness of $M$ and transversality of the cone fields, there are constants $\tau_1,\tau_2>0$ and $K<\infty$ such that, any pair of continuous foliations $\FF^1,\FF^2$ tangent to $C^u_\alpha,C^s_\alpha$ have a product structure with these constants.
We fix $\eta>0$ small enough so that $\Lambda^{1-\eta}e^{-\eta}>1$. $\eta>0$ and $N\geq1$ being fixed, Lemma \ref{lem:deconcentrate-f} yields two numbers $h_0<h_\top(f)$ and $r_0>0$ such that for any $\mu\in\Prob_\erg^{h_0}(f)$, $\mu(B_{r_0})<\eta$. We fix $\gamma>0$ so small that $h_1:=h_0+d\gamma<h_\top(f)$.
We reduce $\eps_0$ so that $\eps_0>0$ and $\eps_0$ is less than the following:
\begin{itemize}
\item $\eps_1(f,\Lambda,\alpha)$,
\item $\eps_2(f,\Lambda,\alpha,\tau_1,\tau_2,K)$,
\item $\tau_2/K_*$,
\item $r_0/(1+2K_*+KK_*)$, and
\item $R_0/KK_*,1/(KK_*+2)^2.$
\end{itemize}
where $\eps_1,\eps_2$ have been defined in Theorem \ref{thm:foliations} and in Proposition \ref{prop:flat}.
\subsection{Entropy decrease under $\pi$}
We show that measures with large entropy for $g$ project to measures with large entropy for $f$.
As $0<\eps<\eps_1$, Theorem~\ref{thm:foliations} yields $g$-invariant center-unstable and center-stable foliations $\FF^{cu},\FF^{cs}$ with $C^1$ leaves. Recall that the shadowing Lemma defines a factor map $\pi:M\rightarrow M$ with
$$\diam(\pi^{-1}(\pi(x)))\leq K_*d_{C^0}(g,f)<K_*\eps < \tau_2.$$
Let $\nu\in\Prob_\erg^{h_1}(g)$. As $0<\eps<\eps_2$, Proposition~\ref{prop:flat} gives a set $X\subset M$ with $\nu(X)=1$ and $\sigma=cu$ or $cs$, such that $\pi^{-1}(\pi(x))\cap X\subset\FF^\sigma_{\tau_2}(x)$ for $\nu$-a.e. $x\in M$. We assume that $\sigma=cu$ and leave the similar case $\sigma=cs$ to the reader.
Let $\mu:=\pi_*(\nu)\in\Prob_\erg(f)$. Recall that inverting a transformation does not change its measure-theoretic entropy so we have the following (easy extension of the) Ledrappier-Walters~\cite{LW} inequality:
$$
h(g^{-1},\nu) \leq h(f^{-1},\mu) + \int_M h_\top(g^{-1},\pi^{-1}(\pi(x))\cap\FF^{cu}(x)) \, \nu(dx).
$$
The dilation under $g^{-1}$ of the center-unstable
leaves is bounded by $e^{\gamma}$, so $h_\top(g^{-1},\FF^{cu}_\delta(x))\leq \dim \FF^{cu}\cdot \gamma$.
It follows that
\begin{equation}\label{eq:loss-entropy}
h(f,\mu) \geq h(g,\nu)-d\gamma > h_0.
\end{equation}
\subsection{Entropy-conjugacy}
We let $0<\eps<\eps_0$ and pick $g\in\Diff^1(M)$ satisfying (H1)-(H3) from Theorem~\ref{thm:main}. Let
$$
M':=\{x\in M:\pi^{-1}(\pi(x))=\{x\} \}\text{ and }M'':=\pi(M').
$$
These are measurable subsets. We show that $M'$ and $M''$, have full measure with respect to any measure in $\Prob_\erg^{h_1}(g)$ and $\Prob_\erg^{h_0}(g)$ respectively, with $h_0,h_1<h_\top(f)\leq h_\top(g)$ defined above.
First, consider $\nu\in\Prob_\erg^{h_1}(g)$.
From \eqref{eq:loss-entropy}, Proposition \ref{lem:deconcentrate-f} yields $\pi_*(\nu)(B_{r_0})<\eta$. But
$$
\pi^{-1}(B_{r_0}) \supset \bigcup_{i=1}^N B(x_i,r_0-2K_*\eps_0)\supset B_{r+KK_*\eps}.
$$
Indeed $r+KK_*\eps\leq \eps+KK_*\eps<r_0-2K_*\eps$. It follows that
$$\nu(B_{r+KK_*\eps})\leq \nu(\pi^{-1}(V_{r_0}))=\mu(V_{r_0})<\eta.$$
Let $x$ be a $\nu$-typical point and let $y\in \pi^{-1}(\pi(x))$. Note that $d(x,y)<K_*\eps<\tau_1$, hence the following points are well-defined: $y^s:=\FF_{\tau_2}^{cs}(x)\cap\FF_{\tau_2}^{cu}(y)$ and $y^u:=\FF_{\tau_2}^{cu}(x)\cap\FF^{cs}_{\tau_2}(y)$. The transversality of $\FF^{cu}$ and $\FF^{cs}$ implies that $d(x,y^s)\leq K d(x,y)\leq KK_*\eps$ and, likewise, $d(x,y^u)\leq KK_*\eps$.
As $g$ is $\gamma$-nearly hyperbolic and $y^u\in\FF^{cu}_{\tau_2}(x)$ we have
$$
d(g(x),g(y^u))\geq e^{-\gamma} d(x,y^u).
$$
As $g$ respects the domination of $f$ and $KK_*\eps<\rho:=(\eps^{1/2}-2\eps)$ we have $g(y^u)-g(x)\in C^u_\alpha(g(x))$.
Consider now the special case where $x\notin B_{r+KK_*\eps}$. Then $d(x,y^u)<KK_*\eps<R_0$ and $y^u-x\in C^u_\alpha(x)$. As $x,y^u\notin B_r$ and $d(x,y^u)<R_0$, we can use the estimates \eqref{eq:exp-Anosov1} and \eqref{eq:exp-Anosov2} and obtain the better lower bound
$$
d(g(x),g(y^u))\geq \Lambda d(x,y^u).
$$
Define $m:M\to\RR$ by $m(x)=\Lambda^{-1}$ if $x\notin B_{r+KK_*\eps}$ and $m(x)=e^{-\gamma}$ otherwise. An induction yields:
$$
\forall n\geq0\quad d(g^n(y^u),g^n(x)) \geq \prod_{k=0}^{n-1} m(g^k(x)) d(x,y^u)
$$
and Birkhoff Ergodic Theorem implies, for $\nu$-a.e. $x\in M$ that
$$
\lim_{n\to\infty} (1/n)\sum_{k=0}^{n-1} \log m(g^kx) \geq (1-\eta)\log\Lambda-\eta\gamma > 0.
$$
As $d(g^n(y^u),g^n(x))\leq KK_*\eps$ for all $n$, we must have $x=y^u$. Likewise $x=y^s$. Thus, $x=y$ a.e. and $\nu(M')=1$.
Let $\mu\in\Prob_\erg^{h_0}(f)$. Proposition \ref{lem:deconcentrate-f} directly shows $\mu(V_{r_0})<\eta$. By compactness, there exists $\nu\in\Prob(g)$ with $\pi_*(\nu)=\mu$ and we can conclude as above that, for $\mu$-a.e. $x$, $\pi^{-1}(x)$ is a single point: $\mu(M'')=1$.
\subsection{Almost expansivity}
In the previous section, we showed that $\pi(y)=\pi(x)$ implies $y=x$ for $\nu$-a.e. $x\in M$ and all $y\in M$ whenever $\nu\in\Prob_\erg^{h_1}(g)$. The hypothesis $\pi(x)=\pi(y)$ was only used to show that $\sup_{n\in\ZZ} d(g^nx,g^ny)< K_*\eps$. Hence, the above reasoning implies that $K_*\eps>0$ is an expansivity constant with respect to all large entropy measures of $g$. This finishes the proof of Theorem \ref{thm:main}.
\section{Proof of Theorem \ref{thm:applyToBV}}\label{s.BV}
In this section we prove that there is a $C^1$-entropically stable, $C^1$-robustly transitive diffeomorphism $g$ of the 4-torus which is not partially hyperbolic. More precisely, we check that the Bonatti-Viana example of a non-partially hyperbolic, robustly transitive diffeomorphism satisfies the assumptions of Theorem \ref{thm:main}. Then, we show that the construction can be modified to obtain arbitrarily large symbolic extension entropy $h_{\operatorname{sex}}(g)$ as stated at the end of Theorem \ref{thm:applyToBV}.
Let $A$ be a $4$ by $4$ matrix with integer entries and determinant one with four distinct real eigenvalues where
$$
0<\lambda_1< \lambda_2 < 1/3<3<\lambda_3< \lambda_4
$$
and that the induced hyperbolic toral automorphism, $f_A$, on the 4-torus has at least 4 fixed points, say $p,q,r,s$.
Following Bonatti and Viana \cite{BV}, one of the fixed points, say $s$, will be left alone to ensure the robust transitivity. A deformation will be done around two others, say $p$, respectively $q$, to forbid hyperbolicity and the existence of any invariant subbundle of the central-stable, respectively central-unstable, subbundle.
The last point $r$ will be used to obtain diffeomorphism with no symbolic extension using techniques from \cite{BD}. We must check that this construction can be performed under the assumptions of Theorem \ref{thm:main} for $N=3$ and $\alpha,\gamma,\eps$ small enough, i.e., smaller than $t(f,N)$ and $\Lambda>\frac{\eps^{1/2}+2\eps}{\eps^{1/2}-2\eps}$.
\begin{figure}[htb]
\begin{center}
\psfrag{A}{$f_A$}
\psfrag{B}{$$}
\psfrag{q}{$q$}
\psfrag{r}{$q_1$}
\psfrag{s}{$q$}
\psfrag{t}{$q_2$}
\psfrag{u}{$q_1$}
\psfrag{v}{$q$}
\psfrag{w}{$q_2$}
\psfrag{C}{$f_0$}
\includegraphics[width=5in]{bv.eps}
\caption{Bonatti Viana construction}\label{f.bv}
\end{center}
\end{figure}
We begin by deforming $f_A$ around the two fixed points $p,q$. Fix $\eta>0$ so small that $\lambda_3-\eta>1$. Let $\gamma\in(0,t)$ small enough so that
$$
(\lambda_3-\eta)/e^\gamma>1.
$$
We fix $\epsilon\in (0,t)$ such that the balls of radius $2\epsilon$ around $p,q,r,s$ are disjoint and
$$1<\frac{\epsilon^{1/2} +2\epsilon}{\epsilon^{1/2} -2\epsilon}<\frac{\lambda_3-\eta}{e^{\gamma}}.$$
We deform $f_A$ into $f_0$ inside $B(q,\epsilon/2)$
keeping $\mathcal F_A^u$ invariant. We do this in two steps.
In the first step, we do a pitchfork bifurcation around $q$ in the stable direction $\lambda_2$.
The stable index of $q$ changes from 2 to 1 and two new fixed points $q_1$ and $q_2$ are created. Then we perturb the diffeomorphism in a neighborhood of $q_2$ so that the contracting eigenvalues become complex; see Figure~\ref{f.bv}.
To be more precise, let $D^2$ be the two dimensional disk and $\phi:D^2\times D^2\rightarrow \mathbb{T}^4$ be a linear chart mapping
\begin{itemize}
\item 0 to $q$,
\item disks $D^2\times\{y\}$, $D_xg_{s_1}^{-1}(C^{cs}_\beta)\subset C^{cs}_\beta$ into the stable leaves of $f_A$, and
\item disks $\{x\}\times D^2$, into the unstable leaves of $f_A$.
\end{itemize}
Let $\chi:D^2\rightarrow [0,1]$ be a smooth cutoff function: so $\chi(0)=1$ and $\chi$ is 0 in a neighborhood of the boundary of $D^2$.
Let $\Psi$ be a volume preserving vector field on $D^2$ such that $\Psi$ has a saddle singularity at the origin with one axis, $e_2$, being expanding and the other, $e_1$, contracting,and $\Psi$ is zero in a neighborhood of the boundary of $D^2$. Let
$
\tilde{\Psi}(x,y)=(\chi(y)\Psi(x), 0).
$
We denote by $\phi_*\tilde{\Psi}$ the push-forward and by $(\Psi)_a$ the time $a$ of the flow defined by a vector field $\Psi$. Let
$$
f_{A,a}=(\phi_*\tilde{\Psi})_a\circ f_A
$$
Observe that the point $q$ remains fixed for all $a\geq 0$ and that the weakest contracting eigenvalue, $\lambda_2=\lambda_2(q,a)$, of $Df_{A,a}(q)$ increases as $a$ increases. It is easy to arrange it so that the expansion at other points is not stronger than that at $q$. So there exists some $a_0>0$ such that the eigenvalue in the direction $e_2$ is 1 for $Df_{A,a_0}(q)$. For $a>a_0$ we have expansion in this $e_2$ direction. Fix $a_1$ larger, but sufficiently close, to $a_0$ such that $\lambda_2(q,a_1)\leq e^{\gamma/2}$. Note that $f_{A,a_1}$ is $\gamma/2$-nearly hyperbolic.
We let $g_0=f_{A,a_1}$ and perturb $g_0$ in a neighborhood of $q_2$ that is disjoint from $q$, using a similar, smaller chart. Let $\Phi$ be a volume preserving vector field of $D^2$ that is zero in a neighborhood of the boundary of $D^2$ and defines a fixed point of center type at the origin. Let
$
\tilde{\Phi}(x,y)=(\chi(y)\Phi(x), 0)
$
and
$$
g_b=(\phi_*\tilde{\Phi})_b\circ g_0.
$$
For some $b_0>0$, the two contracting eigenvalues of $q_2$ for $Dg_{b_0}(q_2)$ become equal. For $b_1$ slightly larger, these eigenvalues are (non-real) complex conjugates.
Note also that the creation of fixed points with different indices prevents
the topologically transitive map from being Anosov.
These non-real eigenvalues also forbid the existence of a one-dimensional invariant sub-bundle inside $E^{cs}$.
The differential of $g_{b_1}$ at each point of $M$ has the following form, using block matrices, in the eigenbasis $(v_1,v_2,v_3,v_4)$ (which we can and do assume to be orthonormal)
$$
\left(\begin{matrix}
\Lambda_{cs} & K \\
0 & \Lambda_u
\end{matrix}\right) \text{ where }
\Lambda_u = \left(\begin{matrix}
\lambda_3 & 0 \\
0 & \lambda_4
\end{matrix}\right)
$$
and $\Lambda_{cs}$ and $K$ are (variable) $2$-by-$2$ matrices with
$$
\|\Lambda_{cs}\|:=\sup_{x\in M}\sup_{\|v\|=1} \|\Lambda_{cs}(v) \|\leq e^{\gamma/2}.
$$
The stable foliation for $f_A$ is invariant under $g_{b_1}$, even though its tangent vectors are not necessarily contracted under $Dg_{b_1}$. Thus, any thin cone field $C^s_\alpha$ will be invariant under $g_{b_1}^{-1}$.
More specifically, the inverse of the above matrix is
$$\left(\begin{matrix}
\Lambda_{cs}^{-1} & -\Lambda_{cs}^{-1}K\Lambda_u^{-1} \\
0 & \Lambda_u^{-1}
\end{matrix}\right).
$$
On the one hand, fixing $\alpha\in (0,t)$ small enough so that
\begin{equation}\label{eq:invcon-cs1}
\alpha < \frac{\lambda_3-e^{\gamma/2}}{\|K\|}
\end{equation}
ensures the invariance $D_xg_{s_1}^{-1}(C^{s}_\alpha)\subset C^{cs}_\alpha$
and such that for all non-zero vectors $v\in C^s_{\alpha}$ and $w\in C^u_{\alpha}$,
$$\|Df v\| < e^\gamma\|v\| \text{ and } \|Dfw\|> (\lambda_3-\eta)\|w\|.
$$
Recall that the cones are defined using the invariant splitting of the original map $f_A$.
On the other hand, the vectors in the unstable subbundle for $f_A$ are still expanded by $Dg_{b_1}$, but the subbundle is no longer invariant.
Let
$C$ be the complement of the center-stable conefield, i.e., $C(x):=\overline{T_xM\setminus C^s_\alpha(x)}$. Then $C$ is an invariant strong-unstable cone field for $g_{s_1}$, but usually very wide. To rectify this, we modify $g_{b_1}$ in $B(q, \epsilon)$.
One defines $f_1$ around $p$ and $q$ by
$$f_1:=L\circ g_{s_1}\circ L^{-1}$$
where
$$
L = \left(\begin{matrix}
\alpha^2 & 0 & 0 &0\\
0 & \alpha^2 & 0 & 0\\
0 & 0 & 1 & 0 \\
0 & 0 & 0 &1
\end{matrix}\right)
$$
We set $f_1=g_{s_1}$ elsewhere. This yields a diffeomorphism since $f_A=L\circ f_A\circ L^{-1}$. $C^{u}_{\alpha}$ is mapped to $C$ by $L^{-1}$, so is an invariant cone field for $f_1$. Also $L^{-1}(C^s_\alpha)\subset C^s_\alpha$. So $f_1$ preserves the two cone fields $C^s_\alpha$ and $C^u_\alpha$ and is $\gamma$-nearly hyperbolic.
We explain why $f_1$ ($\alpha, \rho,\Lambda)$-respects the domination for $f_A$ where $\rho=(\epsilon^{1/2}-2\epsilon)$ and $\Lambda=(\lambda_3-\eta )/e^\gamma$.
The cones $C^u_\alpha(x)$ and $C^s_\alpha(x)$ are constant. Hence, for all $x\in M$ and $y,z\in B(x, \rho)$, if $y-x\in C^u_\alpha(x)$ and $f_1(z)-f_1(x)\in C^s_{\alpha}(f_1(x))$, then $f_1(y)-f_1(x)\in C^u_{\alpha}(f_1(x))$ and $z-x\in C^s_{\alpha}(x)$. Moreover,
$$\frac{\|f_1(y)-f_1(x)\|}{\|f_1(z)-f_1(x)\|}\geq \frac{(\lambda_3-\eta)(\|y-x\|}{e^\gamma\|z-x\|}.$$
Hence,
$$\frac{\|f_1(y)-f_1(x)\|}{\|y-x\|}>\frac{\lambda_3-\eta}{e^\gamma}\frac{\|f_1(z)-f_1(x)\|}{\|z-x\|}=
\Lambda\frac{\|f_1(z)-f_1(x)\|}{\|z-x\|}.$$
Hence, the map $f_1$ ($\alpha, \rho,\Lambda)$-respects the domination for $f_A$. $f_1$ is clearly an $(\eps,3)$-sparse deformation of $f_A$ and we noticed that it is $\gamma$-nearly hyperbolic.
To finish the construction we repeat the deformation just made on $f_A$ near $p$, on $f_1^{-1}$ in the neighborhood of radius $\epsilon$ around $q$. We get a map $f$ which is robustly transitive, not partially hyperbolic, and has a dominated splitting $T\mathbb T^4=E^{cs}\oplus E^{cu}$ with
$\dim E^{cs}=\dim E^{cu}=2$ (see~\cite{BV} for proofs of these facts). Furthermore, by construction the map $f$ satisfies the hypothesis of Theorem~\ref{thm:main} and so is entropy conjugate to $f_A$ and $C^1$-entropically stable, proving the first half of Theorem~\ref{thm:applyToBV}.
\section{Symbolic Extensions}\label{s.symbext}
We modify the diffeomorphism $f$ constructed in Sec.~\ref{s.BV} so that there is no symbolic extension, as stated in Theorem \ref{thm:applyToBV}. The deformation will be done around the fourth fixed point, $r$, for the diffeomorphism $f$. The construction will closely follow the methods in \cite{DN} and some of the discussion in \cite{DF}.
\begin{figure}[htb]
\begin{center}
\psfrag{A}{$f$}
\psfrag{B}{$$}
\psfrag{q}{$r$}
\psfrag{s}{$r$}
\psfrag{t}{$r$}
\psfrag{C}{$g_0$}
\includegraphics[width=5in]{tangency3.eps}
\caption{Homoclinic tangency for $r$}\label{f.tangency}
\end{center}
\end{figure}
First we fix $\epsilon>0$ small enough for Theorem \ref{thm:main} and modify $f$ in the ball of radius $\epsilon$ centered at $r$, along the center-stable direction, just as in the first step of the deformation about $p$. However,
we pick the parameter $a_1$ so that the differential becomes the identity along the center-stable direction at $r$. We further modify $f$ so that, not only the differential, but the map itself restricted to the center-stable leaf of $r$ is the identity in a small ball $B(r,\tau)$. Now we perturb to obtain a new map $g_0$ such that $r$ is a saddle fixed point in the stable direction with directions that are slightly expanding and contracting (without violating the $\gamma$-near hyperbolicity) and such that $r$ has a homoclinic tangency inside $B(r, \tau)$ in the stable leaf.
See Figure \ref{f.tangency}. As in the previous arguments we can do this in such a way that the deformed map $g_0$ will satisfy the conditions of Theorem \ref{thm:main}.
\begin{figure}[htb]
\begin{center}
\psfrag{D}{$g_1$}
\psfrag{B}{$$}
\psfrag{q}{$r$}
\psfrag{s}{$r$}
\psfrag{v}{$r$}
\psfrag{C}{$g_0$}
\includegraphics[width=5in]{tangency2.eps}
\caption{Creation of a horseshoe}\label{f.tangency2}
\end{center}
\end{figure}
We now perturb $g_0$ to obtain a map $g$ with a neighborhood $\mathcal{V}\subset \mathrm{Diff}^1(\mathbb{T}^4)$ and a $C^1$-residual set $\mathcal{D}\subset \mathcal{V}$ such that each $\tilde{g}\in \mathcal{D}$ has no symbolic extension.
We first perturb in a $C^1$ small, but $C^2$ large manner.
The idea is to create a number of transverse intersections near the pervious homoclinic point. See Figure~\ref{f.tangency2}. From these transverse homoclinic points we obtain a locally maximal hyperbolic set with topological entropy larger than some constant. This is now the situation examined in Downarowicz and Newhouse~\cite{DN} where they show the nonexistence of symbolic extensions. The only difference is that we are working on a 2-dimensional leaf of a foliation whereas they are dealing with surfaces. A detailed explanation of this procedure is given in~\cite{DF}. So there exists an open set $\mathcal{V}$ in $\mathrm{Diff}^1(\mathbb{T}^4)$ such that each $g\in \mathcal{V}$ is robustly transitive, not partially hyperbolic and entropically conjugate to $f_A$. Furthermore, there is a $C^1$-residual set $\mathcal{D}$ in $\mathcal{V}$ such that each diffeomorphism in $\mathcal{D}$ has no symbolic extension. This concludes the proof of Theorem~\ref{thm:applyToBV}.
|
\section{Introduction}
The binary system PSR~B1259$-$63/LS~2883~comprises a young radio pulsar with a period 47.8~ms and a Be star LS~2883. With an eccentric (e$\sim$0.87) orbit, the pulsar approaches the periastron every 3.4 years~\citep{1259_radio_94}. The high spin-down power of PSR~B1259$-$63~\citep[$\sim$8$\times$10$^{35}$~erg~s$^{-1}$;][]{Man95} suffices to generate a relativistic pulsar wind (PW). The system is highly variable over an orbital period in radio~\citep[e.g.,][]{1259_radio_05}, X-rays~\citep[e.g.,][]{Hirayama99,Chernyakova_06,Chernyakova_09}, and TeV $\gamma$-rays~\citep{hess_1259_05,hess_1259_09}. The broadband electromagnetic spectrum is believed to result from the interaction of the PW and the stellar wind of LS~2883, the latter being composed of a polar wind and a dense equatorial circumstellar disk. The stellar disk is inclined with respect to the orbital plane~\citep{wex_tilted_98,1259_radio_99} such that the pulsar passes through the disk shortly before and shortly after the periastron passage.
The X-ray flux changes significantly with orbital phase. The 1--10~keV flux increases from $\sim$10$^{-12}~\mathrm{erg}~\mathrm{cm}^{-2}~\mathrm{s}^{-1}$ at apastron to more than $\sim$10$^{-11}~\mathrm{erg}~\mathrm{cm}^{-2}~\mathrm{s}^{-1}$ shortly before and shortly after the periastron passage~\citep[][and references therein]{Chernyakova_09}. The X-ray photon index ($\Gamma_X$) also shows variability related to the orbital period and a spectral state with the hardest spectra ($\Gamma_X\sim1.2$) occurred around the same time as the observed rapid growth of the X-ray flux~\citep{Chernyakova_06}. The unpulsed radio flux increases through the past four periastron passages with respect to other orbital phases, although the light curves show slightly different behaviors during each passage~\citep[][and references therein]{1259_radio_05}.
The first detection of PSR~B1259$-$63/LS~2883~in $\gamma$-rays was made by the H.E.S.S. Cherenkov array through the 2004 periastron passage~\citep{hess_1259_05}, and subsequently in 2007~\citep{hess_1259_09}. The 2004 and 2007 data show that the TeV emission peaks $\sim$10 days before and $\sim$20 days after the periastron passages and the possible dips seen in the TeV light curves before and after the periastron seem to coincide with the stellar disk passage~\cite[see Fig.~1 in][as well as Fig.~\ref{main_plot}]{Kerschhaggl_11}.
An upper limit of $\sim$10$^{-10}$~erg~cm$^{-2}$~s$^{-1}$ was derived using EGRET data~\citep{3rd_egret_cat} for the average $\gamma$-ray emission. A claim made by the AGILE collaboration on the detection of PSR~B1259$-$63/LS~2883~above 100~MeV in a 2-day period in August 2010~\citep{AGILE_1259} has not been confirmed by data collected using the more sensitive Fermi/LAT detector during the same period~\citep{LAT_1259_August}.
During a regular monitoring of $\gamma$-ray emission from PSR~B1259$-$63~shortly before the periastron passage in mid-December, we found the first evidence ($\sim$4$\sigma$) for $\gamma$-ray emission from PSR~B1259$-$63/LS~2883~during a 3-day time interval~\citep{Tam_1259_atel}. The discovery was later confirmed by \citet{Abdo_2nd_tel} who used data collected over 30 days from 2010 November 18.
In this Letter, detailed Fermi analysis results of PSR~B1259$-$63/LS~2883~from 2008 through early 2011 are presented.
\section{Fermi/LAT observations and results}
The Large Area Telescope (LAT) aboard the Fermi Gamma-ray Space Telescope can detect $\gamma$-rays with energies between $\sim$20~MeV and $>$300~GeV~\citep{lat_technical}. The $\gamma$-ray data used in this work were obtained between 2008 August 4 and 2011 February 28. These data are available at the Fermi Science Support Center\footnote{\url{http://fermi.gsfc.nasa.gov/ssc/}}. Due to the discovery of $\gamma$-rays~from PSR~B1259$-$63/LS~2883, a modified sky survey was commenced from 2010 December 27 for 10 days. In this mode the southern hemisphere receives 30\% extra exposure. We used the Fermi Science Tools v9r18p6 package to reduce and analyze the data in the vicinity of PSR~B1259$-$63/LS~2883. Only events that are classified as the ``diffuse'' class or the ``data-clean'' class were used. To reduce the contamination from Earth albedo $\gamma$-rays, we excluded events with zenith angles greater than 105$^\circ$. The instrument response functions ``P6\_V3\_DIFFUSE'' were used.
We carried out unbinned maximum-likelihood analyzes (\emph{gtlike}) of the circular region with a 15$^\circ$ radius centered on the $\gamma$-ray~position of PSR~B1259$-$63~(see below). We subtracted the background contribution by including the Galactic diffuse model (gll\_iem\_v02.fit) and the isotropic background (isotropic\_iem\_v02.txt), as well as all sources in the first Fermi/LAT catalog~\citep[1FGL;][]{lat_1st_cat} within the circular region of 25$^\circ$ radius centered on the $\gamma$-ray~position of PSR~B1259$-$63. We assumed single power laws for all 1FGL sources considered, except for $\gamma$-ray pulsars of which the spectra follow power laws with exponential cut-off~\citep{lat_1st_psr_cat}. The spectral parameter values for sources within 10$^\circ$ from PSR~B1259$-$63~as well as the normalization parameters of the diffuse components were set free.
No significant emission was found before 2010 November 11. We derived upper limits of $\gamma$-ray flux from PSR~B1259$-$63/LS~2883~using data obtained between 2008 August 4 and 2010 November 10 (see Fig.~\ref{main_plot}). However, the source became active roughly when PSR~B1259$-$63~entered the stellar disk~\citep{Tam_1259_atel,Abdo_2nd_tel}.
We first analyzed 0.2--100~GeV data from 2010 November 11 to 2011 February 28. Using a power-law description of PSR~B1259$-$63, the maximized \emph{test-statistic} (TS) value~\citep{Mattox_96} obtained for the pulsar position is 184, corresponding to a detection significance of 13.6$\sigma$. This allows us to classify PSR~B1259$-$63/LS~2883~as a new GeV $\gamma$-ray~source. The best-fit position of the $\gamma$-ray emission is estimated by \emph{gtfindsrc} to be at right ascension (J2000) $=$ 195$\fdg$67 and declination (J2000) $=$ $-$63$\fdg$73 with statistical uncertainty 0$\fdg$06 (0$\fdg$14) at the 68\%(95\%) confidence level, which is consistent with the position of PSR~B1259$-$63. The systematic uncertainty is estimated to be up to $\sim$40\%~\citep{bsl_lat}.
\subsection{Light curve}
We then derived a 0.2--100~GeV light curve composed mostly of 3-day bins (Fig.~\ref{main_plot}). Such bin size ensures enough photon statistics in each bin without loosing information on short time-scale variability. Finer binning (e.g. 1-day) results in low significance for most of the days and high statistical uncertainties in deriving fluxes and photon indices, except for the flaring period that occur between 2011 January 14 and February 3. For data taken more than one month before and two months after the 2010 periastron passage, larger bins were employed for better visualization. From 2010 December 15 to 2011 January 13, no evidence for $\gamma$-ray~emission was found, as first noted in~\citet{Kong_1259_atel}. We therefore derive an upper limit for the above quiescent period. Data points represent the flux values with TS$>$5, for which photon indices are also shown. Otherwise, an upper limit is derived.
As shown in Fig.~\ref{main_plot}, the $\gamma$-ray light curve is highly variable through the 2010 periastron passage: (1) The source started to be active in $\gamma$-rays~about a month before 2010 periastron passage (P1); (2) It remains undetected for about one month since mid-December (Q1); (3) Subsequently, a major flaring period was identified; it peaks at $\sim$35 days after periastron. Having an average flux higher than that in P1 by an order of magnitude, this flare lasted for only $\sim$7 days (P2); (4) A second flare that peaks at $\sim$46 days after periastron, however, lasted longer, so that the source was detected until end of February (P3 and P4). To probe shorter time-scale variability during the flaring period (i.e., P2 and P3), we also produced the light curve with 1-day binning during January~14 to February~4 (Fig.~\ref{zoomin_plot}). Here 0.1--100~GeV photons were used to increase photon statistics. To better demonstrate the variability, we plot data points with TS$>$5 (rather than the more standard criterium of TS$>$9) as flux values since a substantial number of data are of TS between 5 and 9. It can be seen that $\gamma$-rays~from PSR~B1259$-$63/LS~2883~undergo rapid variations down to time scale of one day. Such variability is one of the fastest detected from any Galactic GeV source on the sky. No emission was detected during January 22--26, indicating that the source underwent a short quiescent period between the two major flares.
We also show in Fig.~\ref{main_plot} the orbital X-ray and TeV light curves. The GeV orbital light curve is different than the X-ray and TeV light curves, suggesting that the origin of GeV $\gamma$-rays~is different than the others.
\subsection{Spectral analysis}
As the derived photon indices change significantly with time, we defined four periods that are indicated in Figure~\ref{main_plot} and Table~\ref{4P}. We further divided the 0.1-300~GeV $\gamma$-rays arriving during P1, P2, P3, and P4, respectively, into six energy bins of logarithmically equal bandwidths and reconstructed the flux using \emph{gtlike} for each band independently. A power-law (PL) model for each bin was assumed and the photon spectral index was fixed at a representative value $\Gamma_\gamma=$2.8. While the emission is detected (i.e., TS$>$5) from 1.4~GeV to 20~GeV for P1, no $\gamma$-ray source was apparent at the PSR~B1259$-$63~position in the four bins above 1.4~GeV (the derived TS values $<$5) in the likelihood analysis for P2, P3, and P4, indicating a cut-off at energy $\sim$1~GeV during the flaring period. See Fig.~\ref{SED} for the spectrum derived from a combined analysis of P2 and P3. We therefore attempted to fit the 0.1--100~GeV spectrum with a PL with an exponential cut-off (PLE), as well as a broken PL. We found that PLE describes the spectrum even better during the periods P2, P3, and P4, by $\Delta$TS$=$TS$_\mathrm{PLE}-$TS$_\mathrm{PL}\ge$8, i.e., $\ga$3$\sigma$ in significance levels. The best-fit parameters are shown in Table~\ref{4P}. In particular, the cut-off energies were found to be 310$\pm$160~MeV, 550$\pm$330~MeV, and 250$\pm$95~MeV during the periods P2, P3, and P4, respectively. On the other hand, the parameters provided by the broken PL with all four parameters being free are not well constrained; we therefore do not consider this model.
At a distance 2.3~kpc~\citep{Neg11}, the average energy flux during the flares of $\sim$3$\times10^{-10}$~erg~cm$^{-2}$~s$^{-1}$ corresponds to the $\gamma$-ray~luminosity 1.9$\times10^{35}$~erg~s$^{-1}$. Given the pulsar spin-down luminosity $\sim$8$\times$10$^{35}$~erg~s$^{-1}$, the average $\gamma$-ray~efficiency is about 25\% during the flares.
To demonstrate that GeV emission originates from a component that is temporally and spectrally different than the TeV emission, we put together the 100~MeV to 300~GeV spectra and the $>$300~GeV $\gamma$-ray~spectra obtained for different orbital phases in Figure~\ref{SED}. It is clear that the GeV emission evolves differently compared to the TeV emission, assuming that the TeV behavior does not change dramatically between 2004 and 2010 periastron passages.
\subsection{Correlation between flux and photon index}
In Fig.~\ref{correlation_plot} we plot $\gamma$-ray~flux versus photon index, showing a correlation between these two quantities. We carried out a nonparametric correlation analysis. The computed Spearman rank correlation coefficient between two quantities is $-0.7363$. The probability that this coefficient is different than zero is 0.9959. We have also calculated the linear correlation coefficient (i.e., Pearson's r).
This results in Pearson's r$=-0.7874$ and the probability that this coefficient is different than zero is 0.9986.
\section{Discussion}
High-energy emissions from $\gamma$-ray
binaries (PSR~B1259-63, LS~5039, and LS~I+$61^{\circ}303$) have been discussed using
leptonic models (e.g., Tavani \& Arons, 1997; Dubus 2006; Takata \& Taam 2009)
and hadronic models (e.g., Kawachi et al. 2004; Chernyakova et al. 2009).
In leptonic models, PW particles (electrons and positrons)
are accelerated at the shock where the dynamical
pressure of the PW and that of the stellar wind are
in balance. These particles in turn emit non-thermal photons over a wide range
of energies via synchrotron radiation (radio to GeV $\gamma$-rays) and
inverse-Compton (IC) upscattering off star light
($>$10~GeV $\gamma$-rays), resulting in two peaks in the broadband spectrum.
In hadronic models~\citep[e.g.,][]{Chernyakova_06,Chernyakova_09}, the inverse Compton emission of high-energy electrons resulting from $\pi^0$-decay is responsible for emission from optical up to TeV energies. In this case, the broadband spectrum is rather flat in the energy range of 0.1--100~GeV.
During the pre-periastron epoch, GeV emission at a flux level $\sim6\times10^{-11}$~erg~cm$^{-2}$~s$^{-1}$ is detected from mid-November through mid-December. The fitted spectrum during this period (P1) is relatively hard ($\Gamma_\gamma\sim2$). Such hardness may be consistent with the predictions of both the hadronic and the leptonic models (see below for a discussion of the leptonic interpretation during this period).
The onset of the flare-like
GeV emission at the true anomaly $110^{\circ}-130^{\circ}$ occurred close to
the disk passage at the true anomaly $80^{\circ}-140^{\circ}$, suggesting
that the origin of the $\gamma$-rays~might be related to the disk passage.
On one hand, the observed cut-off at several hundred MeV does not favor hadronic models.
On the other hand, it may be difficult to explain the
flare-like GeV emission using simple leptonic models as well.
First, leptonic models predict a cut-off around 100~MeV for a synchrotron spectrum~\citep[e.g.,][]{Takata09}. Second, the stellar disk pushes the shock towards the pulsar when the pulsar encounters the disk. Although the increase of the magnetic field enhances
the synchrotron power, it also reduces
the synchrotron cooling time of the accelerated particles.
As GeV photons are emitted by fast-cooling particles, these two effects should compensate
each other. Consequently, the shift of the shock position due to
the stellar disk cannot enhance the GeV emission to the level that we observe.
Doppler boosting may provide a plausible mechanism
to produce the GeV flares during January~14 -- February~3, 2011~\citep[see also][]{Kong_sw_11}.
Numerical simulations in the hydrodynamic limit imply that the post
shock bulk flow for the binary system can be accelerated into relativistic
regime because of a rapid expansion of the flow
in the downstream region (Bogovalov et al. 2008). For example,
Dubus et al. (2010) discussed the effects of Doppler
boosting on the orbital modulations of the X-ray and TeV emission.
If the bulk flow is relativistic and oriented
radially away from the star, the Doppler effect may be at work close to the true anomaly corresponding to the direction
of the Earth ($\sim 130^{\circ}$). This is approximately the time when
the GeV flares were observed (i.e., $110^{\circ}-130^{\circ}$). The Doppler effect increases the photon energy
as $E={\cal D}E'$ and the intensity as $I_{\nu}\propto{\cal D}^{3+\alpha}I'_{\nu}$, where ${\cal D}$ is the Doppler factor, and
$\alpha=\Gamma-1$ the spectral index. Non-prime and prime notations represent quantities
in the observer and co-moving frames, respectively.
In $\gamma$-rays, $\alpha_\gamma\sim1-2$ is expected from synchrotron radiation models, which is consistent with the Fermi results.
An enhancement factor of 5--10 in flux during flares (P2 and P3) compared to the emission before periastron (P1; see table~\ref{4P} for details) suggests ${\cal D}\sim$1.5--2.
Suzaku observations indicate a low energy break, i.e., around 10~keV~\citep{Suzaku09}.
Because the photon index of the synchrotron spectrum below the break is
$\alpha=-1/3$, the enhancement factor becomes $D^{3-1/3}\sim3$. Thus, the enhancement in X-rays is
suppressed compared to $\gamma$-rays. It is important to obtain simultaneous observations in X-rays and
gamma-rays during the flaring period to test the boosting model.
In leptonic models, the correlation between $\Gamma_\gamma$ and flux (Fig.~\ref{correlation_plot}) may be
understood as follows. If the Doppler effect is responsible for the flares, photons are boosted to higher energies and the observed synchrotron flux is amplified. However, the Doppler boost does not affect much the photon flux of the IC component,
since more scatters between particles and
incident photons would occur in the Klein-Nishina regime.
Consequently, the increase
in the IC flux by the Doppler boosting and the decrease due to the
Klein-Nishina effect will compensate each other. The energy flux of the synchrotron radiation becomes much higher than the
IC radiation, causing the 0.1--100~GeV spectrum to be dominated by the high-energy tail of the synchrotron spectrum. This gives rise to the steep spectrum ($\sim$3; corresponding to flare emission) shown in Fig.~\ref{correlation_plot}.
The emission before periastron (with $\Gamma_\gamma\sim$2)
may be related to emission originating just behind the shock
where the bulk Lorentz factor is $\sim$1.
Some leptonic models (e.g. Takata et al. 2009;
Dubus et al. 2010) have predicted similar energy fluxes of synchrotron
and inverse-Compton radiation just behind the shock, hence $\Gamma_\gamma\sim2$, which is close to the observed value in the power-law fit. Although there is no evidence for two spectral components during the pre-periastron period (see Fig.~\ref{SED}), this possibility cannot be excluded given the relatively low significance detection, i.e., $\sim$5$\sigma$,
during this period.
PSR~B1259$-$63/LS~2883~is the third known binaries with significant detection in GeV, after LS~I+$61^{\circ}303$~\citep{lat_ls61_303} and LS~5039~\citep{lat_ls5039}. It may be intuitive to compare the three $\gamma$-ray
binaries:
\begin{enumerate}
\item For LS~I+61 303 and PSR~B1259-63, the GeV gamma-ray peak occurs after periastron (and before apastron), though not in the same orbital phase. For LS~5039, the gamma-ray emission peaked at periastron;
\item While no orbital phase-related spectral change has been reported in LS~I+$61^{\circ}303$, the ``softer when brighter'' behavior of GeV $\gamma$-rays~has been found for both LS~5039 and PSR~B1259-63;
\item In all three systems, spectral cutoff is observed at GeV energies during at least part of the orbit;
\item In all three systems, the X-ray and TeV light curves are more correlated whereas GeV $\gamma$-rays~always come out in different orbital phases~\citep{ls5039_xray_tev,ls61_303_xray_tev}. This strongly suggests that GeV $\gamma$-rays~originate from components different than the other two energy bands.
\end{enumerate}
PSR~B1259$-$63/LS~2883~is the only system among these three for which the nature of the compact object is certain. Given the similarities of some emission features of these systems, the results presented here should shed light on the GeV radiation mechanism of the other systems.
\acknowledgments
We acknowledge the use of data and software facilities
from the FSSC, managed by the HEASARC at GSFC. JK and KSC are supported by a GRF grant of HK Government under HKU700908P, and AKHK is supported partly by the National Science Council of the Republic of China (Taiwan) through grants NSC99-2112-M-007-004-MY3 and NSC100-2923-M-007-001-MY3. CYH is supported by research fund of Chungnam National University in 2010.
|
\section{Introduction }
\label{intro}
Cosmic Inflation \cite{Guth:1980zm, Linde:1981mu}
has emerged as a leading theory for the early universe and structure formation which is strongly supported by recent observations \cite{Komatsu:2010fb}. In simple models of inflation, the inflaton field is minimally coupled to gravity with potential flat enough to obtain a long enough
period of inflation to solve the flatness and the horizon problems. Simple models of inflation predict almost scale invariant, almost Gaussian and almost adiabatic curvature perturbations which are in very good agreement with observations. However, with the advance of data in coming years, it is expected that a small but non-zero amount of non-Gaussianities can be detected. This can be used to rule out simple models of inflation such as single field chaotic scenarios.
During last decade there have been extensive efforts to embed inflation in string theory, for a review see \cite{HenryTye:2006uv,Cline:2006hu,Burgess:2007pz,McAllister:2007bg,Baumann:2009ni, Mazumdar:2010sa}. Brane inflation is an interesting realization of inflation from string theory \cite{dvali-tye,Alexander:2001ks,collection,Dvali:2001fw}. In its original form, the scenario
contained a pair of D3 and anti D3 branes moving in the Calabi-Yau (CY) compactification. The inflaton field is the radial distance between the pair so in this sense the inflaton field has a geometric interpretation in string theory. Inflation ends when the distance between the brane and anti-brane reaches the string length scale where a tachyon develops in the open strings spectrum stretched between the pair. Inflation ends soon after tachyon formation and the energy stored in branes tensions are released into closed string modes \cite{HenryTye:2006uv}. However, it was soon realized that the
potential between the pair of brane and anti-brane is too steep to allow a long enough period of
slow-roll inflation. To flatten the potential, it was suggested to put the pair of brane and anti-brane inside a warped throat \cite{Klebanov:2000hb, Giddings:2001yu, Dasgupta:1999ss}, where the potential between D3 and $\D$ is warped down as in Randall-Sundrum scenario
\cite{Randall:1999ee, Kachru:2003sx, Firouzjahi:2003zy, Burgess:2004kv, Buchel, Iizuka:2004ct, Firouzjahi:2005dh}. Dirac-Born-Infeld (DBI) inflation \cite{Silverstein:2003hf, Alishahiha:2004eh} (see also \cite{Kehagias:1999vr}) is a specific model of brane inflation where the mobile brane, the inflaton field, is moving ultra relativistically inside an AdS throat. A novel feature of DBI inflation is the production of large non-Gaussianities which has significant observational
implications to constrain the model parameters. This has attracted considerable interests in literature, for a review of DBI inflation and its implications for non-Gaussianities see
\cite{Chen:2006nt, Chen:2004gc, Shandera:2006ax, Bean:2007eh, Bean:2007hc} and the references therein.
In this work we extend the idea of DBI inflation to the background which is not an AdS
throat. More precisely we will consider the case where the mobile brane is moving in a
Lifshitz background. This geometry has attracted considerable attentions recently
in the context of non-relativistic AdS/CFT correspondence where it may
provide a gravity description for Lifshitz fixed point. We note that Lifshitz fixed
points appear when we are dealing with a physical system at critical point
with anisotropic scale invariance.
Actually at critical points the physics is usually described by scale invariant
phenomena. Typically
the scale invariance arises in the conformal group where we have
\be
t\rightarrow \lambda t,\;\;\;\;\;\;x_i\rightarrow \lambda x_i.
\ee
Here $t$ is time and $x_i$'s are spatial directions of the spacetime.
We note, however, that in many physical systems the critical points are governed by
dynamical scaling in which the space and time scale differently. In fact spatially
isotropic scale invariance is characterized by the dynamical exponent $z$
as follows \cite{Hertz:1976zz}
\be\label{Lif}
t\rightarrow \lambda^zt,\;\;\;\;\;\;x_i\rightarrow \lambda x_i.
\ee
The corresponding critical points are known as Lifshitz fixed points.
In light of AdS/CFT correspondence \cite{Maldacena:1997re} it is natural to seek for
gravity duals of Lifshitz fixed points. Indeed the gravity descriptions of
Lifshitz fixed points have been first considered in \cite{Kachru:2008yh} where
a metric invariant under the scaling \eqref{Lif} where introduced.
The corresponding geometry is given by\footnote{Lifshitz metric typically is a solution
of a gravitational theory coupled to gauge fields \cite{Kachru:2008yh}. Lifshitz metric may also be
a solution of pure gravity modified by curvature squared terms \cite{AyonBeato:2010tm}.}
\be\label{BacLif}
d s^2=-\left(\dfrac{r}{L} \right)^{2z} \mathrm{d} t^2+ \left(\dfrac{r}{L} \right)^2 \mathrm{d} \x^2+\left(\dfrac{L}{r} \right)^2 \mathrm{d} r^2
\ee
where $L$ is the curvature radius of the metric and $r$ is the radial coordinate of the
``Lifshitz throat''. The action of the scale transformation \eqref{Lif} on the metric is given by
\be
\label{scaling}
t\rightarrow \lambda^zt,\;\;\;\;\;\;x_i\rightarrow \lambda x_i,\;\;\;\;\;\;
r\rightarrow \lambda^{-1}r.
\ee
As it has been mentioned in \cite{Kachru:2008yh} although the metric is nonsingular,
it is not geodesically complete and in particular an in-falling object into $r=0$ feels
a large tidal force.
Since at $z=1$ the metric reduces to that of AdS, it is interesting to see how
physical models are
affected when we replace an AdS geometry with Lifshitz metric. In particulate in this paper
we would like to study this effect in a cosmological model. Indeed fast moving branes in an
AdS throat may lead to interesting distinctive inflationary models \cite{Silverstein:2003hf, Alishahiha:2004eh}. Therefore
it is natural to look for inflationary models where the brane moves inside a Lifshitz throat.
Since in this case the background has a free parameter, $z$, one may find interesting
observational predictions when we vary $z$.
We note, however, that unlike AdS geometry which can be easily obtained in the
context of string theory\footnote{An AdS throat may be obtained in type II B string theory
compactified on a Calabi-Yau 3-fold by putting large enough D3 branes on top of each other
at a point on the Calabi-Yau. Of course the throats arising from IIB flux compactification
are not AdS at all length scales, but nevertheless can look approximately AdS in some
scale of energy.}, it is far from obvious how to embed a Lifshitz geometry in string theory.
Nonetheless, it is worth mentioning that string theory realizations of Lifshitz geometries
have been studied in \cite{Hartnoll:2009ns} in the context of strange metallic holography,
see also \cite{Balasubramanian:2010uk, Donos:2010tu, Gregory:2010gx, Cassani:2011sv}.
Actually the present work is motivated by this paper where the authors have studied
the dynamics of a probe D-brane in the background \eqref{BacLif}. The action of the
D-brane probe is given by the DBI action.
It is important to mention that due to lack of a rigorous realization of Lifshitz geometries in string theory, one can not treat the model as a top-down approach for inflationary model building. In particular we anticipate that as long as the string embedding
of Lifshitz geometries is absent, our model suffers from two important shortcomings when
compared to standard DBI inflation. First of all we do not know how the model
allows to have the right
coupling between brane and the Ramond-Ramond (RR) $C_{(4)}$ potential which is necessary to get the correct slow-roll limit for the probe brane moving in the background. Secondly one does not know how to obtain the potential term for the inflaton field which is necessary to support inflation. In standard DBI inflation the potential term can originate from the couplings of the mobile D-brane to background RR fluxes, supplemented by any coupling of the D-brane to the compactification degrees of freedom, including quantum generated effects (see\cite{Silverstein:2003hf} for detail discussions on this point). Despite the above mentioned caveats, we consider our treatment as a phenomenological
approach for inflationary model building with some motivations supported from string theory for the form of the potential or couplings required to support inflation.
The rest of paper is organized as follows. In section \ref{D3-back} we present the action
of mobile brane in a Lifshitz background. In section \ref{background-cosmology} we promote this action to a cosmological set up and look into background inflation and the cosmological perturbations. In section \ref{dim-red} we obtain the general four-dimensional
action with arbitrary matter field and metric perturbations and speculate on its cosmological predictions. The discussion and conclusions are summarized in section \ref{summary}. The details of the gravitational dimensional reduction are relegated into
the Appendix.
\section{D3-Branes in Lifshitz background}
\label{D3-back}
In this section we outline our setup of DBI inflation in Lifshitz background. Before we present our setup, we shall briefly review the conventional models of brane inflation, for a review of brane inflation see \cite{HenryTye:2006uv,Cline:2006hu,Burgess:2007pz,McAllister:2007bg,Baumann:2009ni, Mazumdar:2010sa} and references therein. We will heavily borrow the well-developed ideas and techniques in these models into our background.
In conventional models of brane inflation \cite{dvali-tye,Alexander:2001ks,collection,Dvali:2001fw, Kachru:2003sx, Firouzjahi:2003zy, Burgess:2004kv, Buchel, Iizuka:2004ct, Firouzjahi:2005dh}
branes are moving in an AdS background \cite{Klebanov:2000hb, Giddings:2001yu, Dasgupta:1999ss}.
The action of the mobile branes is given by DBI action supplemented by the Chern-Simons term coming from the coupling of the mobile branes to the background fluxes. In the slow-roll models of warped brane inflation, the inflaton field is the distance between a pair of brane and anti-brane. The anti-brane is dynamically attracted towards the bottom of the AdS throat, the IR region, where the mobile D3-brane is moving slowly towards it from the UV region. The inflationary potential is given by the mutual Coloumbic force between
D3 and $\D$-branes. In this picture inflation ends when the distance between D3 and $\D$ becomes at the order of string scale $l_s$ where a tachyon develops in the open string spectrum and the branes are annihilated. The energy released from the tensions of the colliding branes can be thought as the source of reheating. In the DBI brane inflation \cite{Alishahiha:2004eh, Chen:2004gc, Shandera:2006ax, Bean:2007eh, Bean:2007hc, Chen:2006nt} the mobile brane is moving ultra relativistically inside the throat reaching the asymptotic speed limit. Due to non-standard form of kinetic energy from DBI action, the sound speed in cosmological perturbation theory can be much less than unity. This can have interesting observational consequences such as producing significant amount of non-Gaussianities \cite{Chen:2006nt}.
Here we would like to generalize the above picture of DBI inflation into Lifshitz background.
In light of the above picture of brane inflation, here are the key assumptions we make in our analysis. First, we assume that the Lifshitz background can actually be embedded in string theory.
Second, we assume that the concept of D3-branes with the right coupling to Ramond-Ramond
four form potential $C_{(4)}$ actually exists in this picture
As mentioned previously these are non-trivial assumptions which have yet to be justified in a rigorous string theory set up. Besides these two assumptions, there are other assumptions which are common in models of brane inflation. We assume that all complex structure moduli and the volume moduli are stabilized consistently. There can be back-reactions of volume modulus, background fluxes and branes to the mobile branes. In our effective field theory action, these are interpreted as corrections into the inflationary potential. As in conventional DBI inflation we shall take the phenomenological approach and parametrize the potential appropriately for the inflationary analysis. As usual in brane inflation, one can not avoid the issue of fine-tuning on the mass parameters or couplings \cite{ Kachru:2003sx, Firouzjahi:2003zy, Burgess:2004kv, Firouzjahi:2005dh, Baumann:2006th, Burgess:2006cb, Baumann:2007ah, Chen:2008au, Cline:2009pu, Hoi:2008gc}.
On these issues our position is the same as in conventional models of brane inflation.
\subsection{The Action}
In the Lifshitz background, the Lorentz symmetry in higher dimensions is broken with the metric in the following form
\ba
\label{metr}
d s^2=-\left(\dfrac{r}{L} \right)^{2z} \mathrm{d} t^2+ \left(\dfrac{r}{L} \right)^2 \mathrm{d} \x^2+\left(\dfrac{L}{r} \right)^2 \mathrm{d} r^2+L^2 \mathrm{d} \Omega^2 \, .
\ea
Here $\mathrm{d} \Omega^2$ represents the metric in the angular directions which we left unspecified.
$r$ is the radial coordinate in the ``Lifshitz throat'' with the characteristic length scale $L$. This setup is similar to AdS throat with $z=1$ where $L$ represents the AdS length scale of the throat. We keep the parameter $z$ undetermined taking value
in the range $ z \ge 1$\footnote{For the model considered in \cite{Kachru:2008yh} which
supports metrics of Lifshitz form the reality condition on the fluxes requires $z\ge 1$. Although in the present paper we do not consider a specific gravitational model admitting
Lifshitz geometry, we still assume that this is the case too. } .
All models of brane inflation so far focused on the AdS background with $z=1$.
As explained before, we assume that the mobile brane is moving
in this background governed by the standard DBI and Chern-Simons actions. With this assumption, the action of the mobile D3 brane in background (\ref{metr}) is given by
\ba
\label{action}
S=-T_3 \int \mathrm{d}^4 x \left(\dfrac{r}{L} \right)^{3+z} \left(\sqrt{1-\left(\dfrac{L}{r}\right)^{2+2z}\dot{r}^2}-1 \right) \, .
\ea
Here $T_3$ is the tension of the D3-brane and a dot here and below indicates derivative with respect to $t$. The first term in the bracket, containing the square root, is from the DBI part whereas the second term in the bracket originates from the Chern-Simons term. As before, this form of the action is motivated from the action of D3-brane in AdS throat where there is a specific relation between the background RR four form potential and the warp factor of the throat. This
particular relation in AdS construction reflects the BPS or no-force condition of a mobile D3-brane in the background of AdS geometry in the limit where the brane is moving slowly.
We will be mostly interested in the DBI limit of the action (\ref{action}) where the brane
is moving ultra relativistically in the Lifshitz throat and one can not expand the square root
in Eq. (\ref{action}) perturbativly. However, it would be interesting to look into the slow-roll limit of Eq. (\ref{action}) when one can expand the square root perturbativly. In this limit one finds that the action reduces to that of free field with the action
$S= \frac{1}{2} \int d^4 x \, \dot \phi^2$ where the normalized field $\phi$ is defined via
\ba
\label{dif-eq}
\dot{\phi} \equiv \sqrt{T_3} \, \dot{r} \, \left(\dfrac{r}{L} \right)^{\frac{1-z}{2}} \, .
\ea
We note that in the AdS limit with $z=1$ this gives the well-defined normalization that $\phi= \sqrt{T_3} r$. For general value of $z$ the above differential equation can be integrated easily to yield
\ba
\label{r-phi}
\phi =\begin{cases}
\mu_z \left(\dfrac{r}{L} \right)^{\frac{3-z}{2}} & \text{if} \, \, z \neq 3
\\
\\
\mu_3 \ln(\dfrac{r}{L}) & \text{if} \, \, z=3
\end{cases}
\ea
where we have defined the parameters $\mu_z$ and $\mu_3$ via
\ba
\label{muz3}
\mu_3 \equiv \sqrt{T_3} L \quad \quad , \quad \quad
\mu_z \equiv \dfrac{2 \mu_3}{3-z} = \dfrac{2\sqrt{T_3}}{3-z}\, L \, .
\ea
Note that for $z>3$, $\mu_z$ is negative so $\phi<0$. Also for the case of $z=3$, if
we start with $r<L$, then $\phi<0$. However, in our picture that the brane inside the throat is moving from the UV region towards the IR region, then $\phi$ is always decreasing for all value of $z$. Also we see that the relation between the physically normalized field $\phi$ and the
radial coordinate $r$ is distinctly different for the case with scaling $z=3$. \footnote{In general $d$-dimensional space-time, the case $z=d-2$ is special such that $\phi \sim \ln r$ whereas for other values of $z\ge1$ the relation between $\phi$ and $r$ is a power law.}
Having obtained the physically normalized field $\phi$ in the slow-roll limit, the action \eqref{action} can be rewritten in more compact form which can be applicable to the general case when the brane is moving relativistically
\ba
\label{DBI-Lif}
S=-\int \mathrm{d}^4 x f^{-1} \left(\sqrt{1-f \dot{\phi}^2} -1\right) \, ,
\ea
where
\ba
f(\phi)=\begin{cases}
T_3^{-1} \left(\frac{\mu_z}{\phi} \right)^\alpha & \text{if} \, \, z \neq 3
\\
\\
T_3^{-1} e^{-\frac{6 \phi}{\mu_3}} & \text{if} \, \, z=3
\end{cases}
\ea
and
\ba
\alpha \equiv \dfrac{2(3+z)}{3-z} \, .
\ea
Note that in the case of AdS background we have $\alpha=4$ and $f \sim \phi^{-4}$
as expected.
Interestingly, the form of the D3-action given in Eq. (\ref{DBI-Lif}) is identical to the standard form of DBI inflation. This will be a great help in performing the cosmological analysis in next sections and we can borrow many formulae from the standard DBI inflation
in AdS background. However, we note that despite this formal similarity, there are important
physical differences as one varies the scaling parameter $z$.
\section{DBI Lifshitz Cosmology}
\label{background-cosmology}
Having presented our background, we promote it into a cosmological set up. We couple the world volume of the mobile branes to the FRW metric
\ba
ds^2 = -dt^2 + a(t)^2 \mathrm{d} \x^2 \, ,
\ea
where $a(t)$ is the cosmological scale factor at the background isotropic and homogeneous level. In the cosmological background, the action of D3-brane in Eq. (\ref{action}) is transferred into
\ba
\label{act}
S=-\int \mathrm{d}^4x \, a^3 \left[ f^{-1} \left(\sqrt{1-f \dot{\phi}^2}-1 \right)+V(\phi) \right] \, .
\ea
We have added the potential $V(\phi)$ needed for inflation by hand. As in standard brane inflation, there are many corrections to the dynamics of mobile branes in a throat. This includes the back-reactions from Kahler modulus, background fluxes or branes. Even for the case of
a mobile brane in an AdS throat it is a non-trivial task to calculate all these corrections from string theory \cite{Baumann:2006th, Burgess:2006cb, Baumann:2007ah, Chen:2008au}. At the phenomenological level, one may add the potential term to take into account these back-reactions.
We follow the same phenomenological prescription here and absorb the unknown back-reactions into $V(\phi)$. Whether or not these back-reactions provide the right potential
to sustain a long enough period of inflation is the infamous problem of fine-tuning in brane inflation \cite{ Kachru:2003sx, Firouzjahi:2003zy, Burgess:2004kv, Firouzjahi:2005dh, Baumann:2006th, Burgess:2006cb, Baumann:2007ah, Chen:2008au, Cline:2009pu, Hoi:2008gc}. Our position here is the same as in conventional models of brane inflation where
it is assumed that the parameters of potential can be tuned, at least in principle, so one can obtain a successful period of inflation.
The form of potential is not determined either. For the branes moving in a throat, one may
expect a power law potential in term of its radial coordinate, $V \sim r^{n}$ with unknown
number $n$, which can be either fractional or integer \cite{ Baumann:2007ah}. For our model where the brane is moving towards the IR region of the throat we take $n$ to be a positive number. For $z\neq3$, the scaling between $r$ and $\phi$ given in Eq. (\ref{r-phi})
yields $V \sim r^n \sim \phi^{2n/3-z}$, i.e. a power law inflation potential for $\phi$. On the other hand, for the case with $z=3$, Eq. (\ref{r-phi}) indicates that
$V \sim r^n \sim e^{ n \phi/ \mu_3 } $, i.e. an exponential potential for the inflaton field.
These suggest that for our phenomenological investigations we can take the inflaton potential to
have the following form
\ba
V(\phi)=\begin{cases}
V_0 \, \left( \frac{\phi}{\mu_z} \right)^p & \text{if} \, \, z \neq 3
\\
V_0 \, e^{\frac{n \phi}{\mu_3}} & \text{if} \, \, z=3
\end{cases}
\ea
where
\ba
p\equiv \frac{2n}{3-z} \, .
\ea
The unknown parameter $n$ and the energy scale $V_0$ are left undetermined and should be tuned to support long enough period of inflation and satisfy the WMAP constraints.
In our phenomenological treatments, we have assumed that the form of potential is determined
as a function of $r$, such as $ V \sim r^n$ assumed above, and obtained the corresponding form of $V(\phi)$. Instead, we could have chosen to start with a fixed phenomenological potential $V(\phi)$ and read off its corresponding form $V(r)$. However, the latter approach is less natural. In our setup with arbitrary value of $z$, the relation between $r$ and $\phi$ is
non-linear so if we start with $V(\phi)$ then the form of $V(r)$ would be completely different for the cases of $z=3$ and $z \neq 3$. The naturalness in starting with $V(r)$, as we employed above, originates from the fact that to calculate $V(r)$ one has to incorporate the effects of background fluxes and volume moduli of string compactification. In this picture $r$, being a real string theory coordinate, is more natural and the resulted potential from the back-reactions is expected to be a function of $r$ in the form of $V(r)$.
We note that the case $z=1, n=2$ corresponds to conventional model of DBI inflation
with the potential $m^2 \phi^2/2$ which is vastly studied in the literature. In the analysis below, we consider the cosmological predictions of our model for different values of $z$ and $n$.
\subsection{Background Cosmology }
We couple the action (\ref{act}) to the effective four-dimensional gravity and write down the background cosmological equations.
In the analysis below we mostly follow \cite{Shandera:2006ax} in notations and methods.
As usual the Friedmann equation and the energy conservation equation are
\ba
\label{friedmann}
3 H^2 = \dfrac{\rho}{M_P^2} \quad , \quad \dot \rho + 3 H (\rho+p) =0 \, .
\ea
Here $H= \frac{\dot a}{a}$ is the Hubble expansion rate, $\rho$ and $p$ respectively are the energy density and the pressure
\ba
\label{p}
\rho = f^{-1} (\gamma -1)+ V
\quad , \quad
p=f^{-1}( 1- \gamma^{-1}) -V \, ,
\ea
in which $\gamma$ is the so called ``Lorentz factor" defined by
\ba
\gamma \equiv \dfrac{1}{\sqrt{1-f \dot{\phi}^2}} \, .
\ea
The modified Klein-Gordon equation for the inflaton field is
\ba
\label{phi-eq}
\ddot{\phi}+3H\gamma^{-2}\dot{\phi}+\frac{3}{2}\frac{{f'}}{f}\dot{\phi}^2-\frac{{f'}}{f^2}+\gamma^{-3} \left({V'}+\frac{{f'}}{f^2} \right)=0 \, .
\ea
Following \cite{Shandera:2006ax}, one can cast these equations into Hamilton-Jacobi forms which are more suitable for analytical purposes. Since $\phi$ is monotonically decreasing
as time goes by, we can use $\phi$ as the clock and express the physical parameters in terms of
$\phi$. This yields
\ba
\label{HJ}
3 M_P^2 H(\phi)^2 &=& V(\phi)+f^{-1}(\gamma(\phi) -1) \nonumber \\
\gamma(\phi) &=& \sqrt{1+4 M_P^4 f(\phi) H^{'}(\phi)^2} \nonumber \\
\dot{\phi}(\phi) &=& \dfrac{-2 M_P^2 H'}{\gamma(\phi)}
\ea
where $H'= \partial H/\partial \phi$.
We are interested in the limit where the brane is moving ultra relativistically inside the throat with $\gamma \gg 1$. This corresponds to brane ``speed limit'' where
$\dot \phi \simeq -\frac{1}{\sqrt{f}}$. For $z= 3$ this yields
\ba
\label{speed-z=3}
\phi_{speed} \rightarrow - \frac{\mu_3}{3} \ln \left( \frac{3 t}{L} \right) \quad \quad
(z= 3)
\ea
This indicates that the speed limit is reached only logarithmically in time, much slower than the case of AdS background with $z=1$ where $\phi_{speed} \sim \frac{1}{t}$ \cite{Silverstein:2003hf}. Note that the brane is moving towards the IR region where $r<L$ so that $\phi_{speed} <0$ as seen above.
On the other hand, for $z\neq3$, we obtain
\ba
\label{speed-z}
\phi_{speed} \rightarrow \mu_z \left( \frac{z t}{L}
\right)^{(z-3)/2z} \quad \quad
(z\neq 3)
\ea
In particular, with $z=1$ we obtain $\phi_{speed} \sim \frac{1}{t}$ as expected for
standard DBI inflation. Note that for $z>3$, $\phi$ and $\mu_z$ are negative so as time goes by
$| \phi|$ increases. This is indicated by the fact that the speed limit is speeding up with
a positive power of $t$. On the other hand, for $1 \leq z \leq 3$, $\phi$ is positive so
a negative power of $t$ above indicates that $\phi \rightarrow 0^+$. Of course this picture will
terminate if we want to have a graceful exit from inflation. In our picture, we assume that
there is an anti-brane at the bottom of the throat, $r=r_0$. Once the distance between the mobile brane and anti-brane reaches at the order of string length scale then the system becomes tachyonic and inflation ends quickly after brane and anti-brane annihilation. So the speed limit
described above is in the idealistic limit where the Columbic force between brane and anti-brane is neglected and the brane is moving indefinitely towards the bottom of the throat in the absence of anti-branes.
It is also instructive to look into the number of e-folding, $N$. To solve the flatness and the horizon problem we assume $N =60$.
Using $d N = H dt$, one obtains
\ba
\label{N-eq}
N = \int_{\phi_f}^{\phi_i} d \phi \, H(\phi) \sqrt {f(\phi)} \, ,
\ea
where $\phi_i$ and $\phi_f$ respectively represent the initial and the final value of the inflaton field.
In the speed limit one can easily integrate this expression and find $N$ as a function of $\phi_i$ and $\phi_f$.
However, it turns out that the results can be expressed more easily in terms of $r$-coordinate
\ba
\label{N-r}
N=\sqrt{\frac{ V_0 \mu_3^2}{3 M_P^2 T_3}}
\begin{cases}
\frac{2}{n-2z} \left( \frac{r}{L}
\right)^{\frac{n-2z}{2}} \huge {|}_{r_f}^{r_i} & \left( z \neq \frac{n}{2}\right)
\\
\ln \left( \frac{r_i}{r_f}
\right) & \left( z=\frac{n}{2} \right)
\end{cases}
\ea
where $r_i$ and $r_f$, respectively, are the initial and the final values of the mobile brane's coordinates.
Here we pause to address the issue of ending inflation where $N$ is an increasing function of $r_f$ as in the case of $n<2z$ in Eq. (\ref{N-r}). Interestingly this situation indicates that inflation, or a dS solution, is attractor towards the IR region of the throat. As the brane is moving towards the IR region, inflation proceeds indefinitely. However, as mentioned above, inflation ends in this situations when the distance between brane and anti-brane, located at the bottom of throat $r=r_0$, reaches at the order of string scale. Using the metric (\ref{metr}) the physical distance $d$ between the brane and anti-brane is
$d= L \ln \left(r_f/r_0\right)$ where $r_f$ is the final value of the brane position. Setting
$d=l_s$ for $l_s$ being the string length scale results in
\ba
\label{rf}
r_f = r_0 \exp \left( \frac{l_s}{L}
\right) \, .
\ea
For the case $z=3$, this results in
\ba
\label{phif-z=3}
\phi_f = \mu_3 \left[ \frac{l_s}{L} - \ln \left( \frac{L}{r_0}\right) \right] \, .
\ea
Depending on the sign of the term inside the bracket, $\phi_f$ can be either positive or negative.
However, one can easily arrange such that $L/r_0$ is exponentially large as in GKP
construction \cite{Giddings:2001yu} such that $\phi_f$ is typically expected to be negative.
On the other hand, for $z\neq 3$ the value of $\phi_f$ is obtained to be
\ba
\label{phif-z}
\phi_f = \mu_z \exp \left[ \frac{3-z}{2} \left( \frac{l_s}{L} - \ln \left( \frac{L}{r_0}\right) \right) \right] \, .
\ea
As expected, for $L/r_0$ exponentially large, $\phi_f \rightarrow 0^+$ for $1\leq z <3$
whereas for $z>3$ $\phi_f$ decrease towards more negative values.
After reviewing the background cosmology, we turn to perturbations in general DBI Lifshitz inflation.
\section{Dimensional Reduction and Cosmological Perturbations}
\label{dim-red}
Having presented our background inflationary solution, it is time to address the question of cosmological perturbations in this model. As we shall see below, the perturbations in this model is drastically different than the standard DBI inflation model. The reason is that after dimensional reduction from 5D into 4D, the general covariance is lost explicitly in 4D.
The loss of general covariance is explicit both in the field theory sector and in the gravitational sector. This is because we start from a theory in higher dimension which breaks Lorentz invariance explicitly when $z \neq 1$. Having this said, one may worry that the lack of 4D general covariance may destroy the effective FRW cosmology which we have obtained at the homogeneous and isotropic level. Here we verify that this is not a problem for the background cosmology. To see this, let us uplift metric (\ref{BacLif}) into a homogeneous and isotropic FRW background
\ba
\label{FRW-Lif}
d s^2=-N(t)^2 \left(\dfrac{r}{L} \right)^{2z} \mathrm{d} t^2+ \left(\dfrac{r}{L} \right)^2 a(t)^2 \mathrm{d} \x^2+\left(\dfrac{L}{r} \right)^2 \mathrm{d} r^2
\ea
where as usual $N(t)$ is the lapse function added in order to find the Friedmann constraint equation. Calculating the five-dimensional Ricci scalar $^{(5)}R$ from metric (\ref{FRW-Lif}) we have
\ba
\label{R5-1}
^{(5)}R = \left( \frac{L}{r}\right)^{2 z} R_{FRW} -\frac{1}{L^2} \left( 12+ 6 z + 2 z^2
\right)
\ea
where $R_{FRW}$ is the four-dimensional Ricci scalar constructed from the four-dimensional FRW metric
\ba
R_{FRW} \equiv \frac{6}{N^2} \left(\frac{\ddot a}{d} + (\frac{\dot a}{a})^2 - \frac{\dot a}{a} \frac{\dot N}{N} \right) \, .
\ea
Note that the last term in bracket in Eq. (\ref{R5-1}) has no dynamics in terms of the four-dimensional metric and contributes only to the effective cosmological constant which can be canceled by similar terms from other fields (such as the massive gauge fields which we do not consider here). As a result, the four-dimensional gravitational action, $^{(4)}S_{gr}$, is
\ba
\label{Sgr}
^{(4)}S_{gr} \subset \frac{1}{2 \kappa^2}\int d^5 x \sqrt{- G}\, ^{(5)}R =
\frac{M_P^2}{2}
\int d^4 x N a(t)^3 R_{FRW}
\ea
with the identification
\ba
\label{Mp}
M_P^2 \equiv \frac{1}{ \kappa^2} \int_V d r \left(\frac{r}{L}\right)^{2-z} \, ,
\ea
where $\kappa$ is the five-dimensional gravitational coupling and the integration over the compact volume $V$ is supposed to be finite in order to obtain a finite four-dimensional gravitational coupling. As usual, this can be achieved by gluing the Lifshitz throat into the bulk of CY compactification. Eq. (\ref{Sgr}) clearly demonstrates that we recover the standard FRW action for the isotropic and homogeneous cosmology and our results for the background cosmology coupled with the matter sector given in Eq. (\ref{act}) is indeed justified.
Now we would like to uplift metric (\ref{BacLif}) into a generic cosmological background where the effective 4D metric, once the extra $r$-coordinate is integrated out,
is given by $g_{\mu \nu}= \{ g_{00}, g_{0i}, g_{ij} \}$. The natural ansatz is
\ba
\label{Lif-cosmo}
d s^2=g_{00}\left(\dfrac{r}{L} \right)^{2z} \mathrm{d} t^2+ g_{ij}\left(\dfrac{r}{L} \right)^2
\mathrm{d} x^i \mathrm{d} x^j+ 2 g_{0i} \left(\dfrac{r}{L} \right)^{\kappa} \mathrm{d} t\, \mathrm{d} x^i +
\left(\dfrac{L}{r} \right)^2 \mathrm{d} r^2
\ea
Here $g_{00}, g_{ij}$ and $g_{0i}$ are arbitrary functions of 4D space-time coordinates $x^{\mu}$. We have left the arbitrary scaling parameter $\kappa$ in the $0i$ component of the metric. However, demanding that the scaling (\ref{scaling}) still to hold requires that $\kappa = z+1$.
Before we proceed with the perturbations it should be stressed that the ansatz (\ref{Lif-cosmo})
may not be a solution of the five-dimensional Einstein equation. In principle there are other fields, such as a massive gauge field, which should be added into the action in order to support the Lifshitz geometry (\ref{BacLif}). To study the perturbations one should also study the perturbations of these non-gravitational fields.
In our analysis below we shall concentrate only on the gravitational sector, given by the usual Einstein-Hilbert term $\sqrt{-|G_{MN}|} {^{(5)}}R$, and the inflaton perturbations.
The ansatz (\ref{Lif-cosmo}) with only the metric and inflaton perturbations are rich enough to demonstrate the nature of perturbations in our analysis.
\subsection{Matter Action}
Now we obtain the matter sector Lagrangian with the metric perturbations given in Eq. (\ref{Lif-cosmo}). The Lagrangian for the matter sector is obtained by considering a probe brane moving inside the Lifshitz background whose embedding $X^M$ is a general function of 4D
space-time coordinates
\ba
X^M = \left( x^\mu, r( x^\nu) \right)
\ea
where $r( x^\nu) $ indicates the brane position inside the throat.
The induced metric on the brane, $\bar g_{\mu \nu}$, is given by
\ba
\bar g_{\mu \nu} = \frac{\partial X^M}{\partial{x^\mu} } \frac{\partial X^M}{\partial{x^\mu} }
G_{MN} \, ,
\ea
where $G_{MN}$ is the 5D background metric given by Eq. (\ref{BacLif}). One obtains
\ba
\bar g_{00} &=& g_{00} \left(\frac{r}{L}\right)^{2z} + \left(\frac{L}{r}\right)^2 \dot r^2
\nonumber\\
\bar g_{ij} &=& g_{ij} \left(\frac{r}{L}\right)^2 + \left(\frac{L}{r}\right)^2 \partial_i r \partial_j r
\nonumber\\
\bar g_{0i} &=& g_{0i} \left(\frac{r}{L}\right)^{z+1} + \left(\frac{L}{r}\right)^2 \dot r \, \partial_i r \, .
\ea
To obtain the matter sector action we are interested in calculating $\sqrt{- | \bar g_{\mu \nu} |}$. After some long calculations one obtains
\ba
\label{general-metric}
| \bar g_{\mu \nu} | = | g_{\mu \nu} | \left(\frac{r}{L} \right)^{z+3}
\left[ 1+ g^{00} \left(\frac{r}{L}\right)^{-2(z+1)} \dot r^2
+ \left(\frac{r}{L}\right)^{-4} g^{ij} \partial_i r \partial_j r + 2 \left(\frac{r}{L}\right)^{-(z+3)} g^{0i} \dot r \, \partial_i r
\right] \, .
\ea
Note that here $g^{\mu \nu}$ is the inverse of metric $g_{\mu \nu}$ in the usual sense.
This equation has a very interesting structure. We note that in the limit where $z=1$, the 4D theory becomes general covariant in the matter sector as expected. However, for arbitrary $z \neq1$ the general covariance is lost
in the field theory. In order to connect it to our background inflation analysis, we introduce the physical field $\phi$ as before
\ba
\label{dif-eq}
\dot{\phi} \equiv \sqrt{T_3} \, \dot{r} \, \left(\dfrac{r}{L} \right)^{\frac{1-z}{2}} \, .
\ea
Working with $\phi(x^\alpha)$, the DBI action becomes
\ba
-T_3 \sqrt{- | \bar g_{\mu \nu} |} = -\sqrt{- | g_{\mu \nu} |} f(\phi)^{-1} \left[1+
f(\phi) g^{00} \left(\partial_t \phi\right)^2 + h(\phi) g^{i j} \partial_i \phi \partial_j \phi
+ 2 \ell(\phi) g^{0i} \partial_t \phi \partial_i \phi
\right]^{1/2}
\ea
in which for the case $z\neq 3$ we have
\ba
f(\phi) \equiv T_3^{-1} \left(\frac{\mu_z}{\phi} \right)^\alpha \quad , \quad
h(\phi)\equiv T_3^{-1} \left(\frac{\mu_z}{\phi} \right)^{\alpha'} \quad \quad z\neq 3
\ea
with
\ba
\alpha \equiv \dfrac{2(3+z)}{3-z} \quad , \quad
\alpha' \equiv \dfrac{2(5-z)}{3-z} \, .
\ea
On the other hand, for $z=3$ one has
\ba
f(\phi)\equiv T_3^{-1} e^{-\frac{6 \phi}{\mu_3}} \quad , \quad
h(\phi) \equiv T_3^{-1} e^{-\frac{2 \phi}{\mu_3}} \quad , \quad z=3
\ea
Very interestingly, in both cases one has $\ell(\phi) = \sqrt{f(\phi) h(\phi)}$.
Now adding the Chern-Simons part and demanding that the theory reduces to a free field theory in the slow-roll limit the total brane action is
\ba
\label{D3-action}
{\cal{L}}_{D3} = -\sqrt{- | g_{\mu \nu} |} f(\phi)^{-1} \left\{ \left[1+
f(\phi) g^{00} \left(\partial_t \phi\right)^2 + h(\phi) g^{i j} \partial_i \phi \partial_j \phi
+ 2 \ell(\phi) g^{0i} \partial_t \phi \partial_i \phi
\right]^{1/2} -1\right\} \, .
\ea
This is an interesting theory. It clearly shows that the general covariance is lost in the matter sector. One can restore the general covariance only when $z=1$ and $f(\phi)= h(\phi)=\ell(\phi)$.
Now that the matter sector breaks the general covariance explicitly, we expect that the general covariance is also lost in 4D gravitational theory.
\subsection{Gravitational Action}
\label{gr-action}
In this subsection we would like to calculate the four dimensional gravitational action
$\sqrt{- |g_{\mu \nu}|}\, ^{(4)}R$ from the five-dimensional gravitational action $\sqrt{- G}\, ^{(5)}R$ where $G_{MN}$ is given by Eq. (\ref{Lif-cosmo}). To start, from Eq. (\ref{Lif-cosmo}) we have $\sqrt {-|G_{MN}|} = \sqrt{-|g_{\mu \nu}|} \left( \frac{r}{L} \right)^{z+2} $ . It is interesting that the 4D and 5D determinants are proportional to each other. One can check that this is possible if one takes $\kappa = z+1$
as we did.
Now we consider the gravitational dimensional reduction. One key observation is that although
the four-dimensional diffeomorphism is expected to be broken explicitly but the metric (\ref{Lif-cosmo}) is still invariant under the three-dimensional diffeomorphism and time rescaling given by
\ba
\label{3d-diff}
t \rightarrow \tilde t = \tilde t(t) \quad , \quad
x^i \rightarrow \tilde x^i= \tilde x^i (x^j, t) \, .
\ea
As a result the dimensionally reduced gravitational action is expected to be written in such a way that the three-dimensional diffeomorphism invariance is explicit. In general one expects that the result to be expressed in terms of the intrinsic three-dimensional Ricci scalar $^{(3)}R$ and the extrinsic curvature $K_{ij}$ which describes the curvature of the three-dimensional surface $t=constant$ in four dimension.
To find $K_{ij}$ it is better as usual to employ the ADM formalism where the metric
(\ref{Lif-cosmo}) is written as
\ba
\label{ADM}
ds^2 = -(\frac{r}{L})^{2z} N^2 dt^2 + g_{ij} \left[ (\frac{r}{L}) dx^i + N^i (\frac{r}{L})^z dt
\right] \left[ (\frac{r}{L}) dx^j + N^j (\frac{r}{L})^z dt
\right] \, ,
\ea
so
\ba
g_{00} = -N^2 + g_{ij} N^i N^j \quad , \quad g_{ij} = ^{(3)}g_{ij} \quad ,\quad
g_{0i} = g_{ij} N^j \, .
\ea
Note that we use $G_{MN}$ for the 5D metric while $g_{\mu \nu}$
is the 4D metric as defined in (\ref{Lif-cosmo}). One can also check that
\ba
g^{00} = -\frac{1}{N^2} \quad , \quad g^{ij} = ^{(3)} g^{ij} - \frac{N^i N^j}{N^2}
\quad , \quad g^{0i} = \frac{N^i}{N^2}
\ea
It is important to note that we consider a general metric perturbations so
$N=N(t, x, y, z)$, i.e. a ``non-projectable'' metric \cite{Blas:2009qj} which is still consistent
with the time rescaling and the three-dimensional diffeomorphism given in Eq. (\ref{3d-diff}).
The extrinsic curvature is given by
\ba
K_{ij} = \frac{1}{2N} \left( \dot g_{ij} - ^{(3)} {\nabla_i N_j} - ^{(3)} \nabla_j N_i
\right)
\ea
where $^{(3)} \nabla$ represents the three-dimensional covariant derivative constructed from
the metric $^{(3)} g_{ij}$.
We relegate the details of the gravitational dimensional reduction analysis into
Appendix \ref{gdr} and quote the final results. After neglecting terms which are either
total derivatives or contribute only to the effective cosmological constant term such as the last term
in Eq. (\ref{R5-1}) one obtains
\ba
\label{gravity1}
\sqrt {-|G_{MN}|} \, {^{(5)}}R \rightarrow \left(\frac{r}{L}\right)^{2-z} \sqrt {-|g_{\mu\nu}|}
\left[ {^{(4)}}R + \Omega\, {^{(3)}}R
\right] \,
\ea
where
\ba
\label{Omega}
\Omega\equiv (\frac{r}{L})^{2(z-1)} -1 \, .
\ea
In the limit where $z=1$ and $\Omega=0$ we recover the standard four dimensional gravity
which is general covariant. However, for non-zero value of $\Omega$ the 4D general covariance is broken into three-dimensional general covariance. This is also in light with the loss of the 4D general covariance in the matter sector, Eq. (\ref{D3-action}). Note that for the FRW background with ${^{(3)}}R=0$, Eq. (\ref{gravity1}) reproduces Eq. (\ref{R5-1}) as expected. On the other hand noting that $^{(4)}R = ^{(3)}R +K^2 - K_{ij}K^{ij}$,
we can replace $^{(3)}R$ in terms of the extrinsic curvature to obtain
\ba
\label{gravity2}
\sqrt {-|G_{MN}|}\, {^{(5)}}R \rightarrow \left(\frac{r}{L}\right)^{2-z} \sqrt {-|g_{\mu\nu}|}
\left[ \left(1+ \Omega \right) {^{(4)}}R + \Omega \left( K^2 - K_{ij} K^{ij} \right)
\right]
\ea
Now we perform the compactification. Performing the integral over the internal manifold volume $V$ one obtains
\ba
\label{grav-final}
\frac{1}{2 \kappa^2} \int d^5 x \sqrt{-G} ^{(5)}R &&\rightarrow \int d^4 x \, \sqrt{-|g_{\mu \nu}|}\frac{M_P^2}{2} \left[ ^{(4)}R + \bar \Omega\, {^{(3)}}R \right] \nonumber\\
&&= \int d^4 x \, \sqrt{-|g_{\mu \nu}|}\frac{M_P^2}{2} \left[ (1+ \bar \Omega) ^{(4)}R + \bar \Omega\,
\left( K^2 - K_{ij} K^{ij} \right)\right] \, ,
\ea
where $M_P$ is defined as in Eq. (\ref{Mp})
and
\ba
\bar \Omega \equiv \frac{1}{\kappa^2 M_P^2} \int_V dr \, \Omega \, \left(\frac{r}{L}\right)^{2-z} = \frac{1}{\kappa^2 M_P^2} \int_V dr \, \left( \frac{r}{L} \right)^{2-z}
\left[ -1 + \left( \frac{r}{L} \right)^{2(z-1)}
\right] \, .
\ea
Eq. (\ref{grav-final}) is our final result for the gravitational dimensional reduction.
One natural question is the magnitude and the sign of the parameter $\bar \Omega$ which measures the extent of the 4D general covariance breaking. To estimate $\bar \Omega$ one has to know the details of the compactification. As in conventional models of brane inflation suppose the Lifshitz throat is extended between the radial coordinate $r_0 < r<R$ so
$r_0$ is the IR end of the throat while $R$ is the UV region where the Lifshitz throat is assumed to be glued smoothly to the bulk of the CY compactification. Using Eq. (\ref{Mp})
to eliminate $\kappa^2 M_P^2$ we obtain
\ba
\label{omeg-1}
\bar \Omega = -1 + \frac{\int_V dr \left(\frac{r}{L}\right)^z}{\int_V dr\, \left(\frac{r}{L}\right)^{2-z}} \, .
\ea
To calculate $\bar \Omega$ we consider the cases $z<3$, $z=3$ and $z>3$
separately.
For $z < 3$ and assuming that $r_0 \ll R$, we obtain
\ba
\bar \Omega \simeq -1 + \frac{3-z}{z+1} \left(\frac{R}{L}\right)^{2(z-1)}
\quad \quad z < 3 \, .
\ea
Depending on the value of the ratio $R/L$, $\bar \Omega$ can be either positive or negative.
Defining $R_c$ as when $\bar \Omega=0$, we obtain
\ba
R_c = L \left(\frac{3-z}{z+1}\right)^{1/2 (z-1)} \, .
\ea
So for $R<R_c (R>R_c)$ one has $\bar \Omega <0 (\bar\Omega >0)$.
Interestingly at $R=R_c$, $\bar \Omega$ vanishes although we have $z \neq 1$. This means that the 4D theory is general covariant in the gravitational sector if the Lifshitz throat is glued to the CY compactification at the critical radius $R=R_c$. For generic value of $R$ and assuming that $R/L$ typically is order of few, we expect that the magnitude of $\bar \Omega$ to be order of few.
On the other hand for $z>3$, we have
\ba
\bar \Omega \simeq -1 + \frac{z-3}{z+1} \left(\frac{R}{L}\right)^{2(z-1)}
\left(\frac{r_0}{R}\right)^{z-3}
\quad \quad z > 3 \, .
\ea
We know that $r_0/R$ is exponentially small \cite{Giddings:2001yu}, so for values of
$R \lesssim L$ we expect to have $\bar \Omega <0$. However, by choosing $R$ to be
much bigger than $L$ one can tune $\bar \Omega$ to cross zero and become positive too.
Finally for $z=3$, we obtain
\ba
\bar \Omega \simeq -1 + \frac{(\frac{R}{L})^4}{4 \ln(\frac{R}{r_0})} \,
\quad \quad \quad \quad z = 3 \, .
\ea
As for the case
$z<3$, one can find a critical value $R_c$ where $\bar \Omega=0$. Assuming that typically
$ \ln R/r_0 $ is order few, we conclude that $R_c \sim L$.
For $R < R_c (R>R_c)$ we have $\bar \Omega <0 (\bar \Omega >0)$.
Having obtained the action for the matter and the gravitational sectors, Eqs. (\ref{D3-action}) and (\ref{grav-final}), the total action, $S= \int d^4 x { L}_{total}$, therefore is
\ba
\label{S-total}
{ L}_{total} &=& \sqrt{-|g_{\mu \nu}|}\frac{M_P^2}{2} \left[ ^{(4)}R + \bar \Omega\, {^{(3)}}R \right] \nonumber\\
&-&\sqrt{- | g_{\mu \nu} |} f(\phi)^{-1} \left\{ \left[1+
f(\phi) g^{00} \left(\partial_t \phi\right)^2 + h(\phi) g^{i j} \partial_i \phi \partial_j \phi
+ 2 \ell(\phi) g^{0i} \partial_t \phi \partial_i \phi
\right]^{1/2} -1\right\}
\ea
Eq. (\ref{S-total}) indicates an interesting theory where the general covariance is broken explicitly both in the matter and the gravitational sectors. This action is similar to the classifications employed in effective field theory of inflation (EFTI) literature \cite{Cheung:2007st}. In EFTI studies the 4D general covariance is broken because the inflaton field
introduces preferred three-dimensional time foliation slices.
This is also the case for our matter sector where the non-trivial $\phi$-dependences break explicitly the 4D general covariance. However, the lack of 4D covariance in gravitational sector is more non-trivial. It originates from the fact that the higher-dimensional
theory breaks the Lorentz invariance at the level of solution. After compactification, this manifests itself with an extra scalar field degree of freedom. This is reminiscent of the extra scalar degree of freedom in Horava-Lifshitz theory of gravity \cite{Horava:2009uw, Charmousis:2009tc, Li:2009bg, Blas:2009yd, Blas:2009qj}.
Having presented our total action in Eq. (\ref{S-total}) one can look into the cosmological perturbation theory in details and check the observational predictions. Specifically the model seems to have interesting predictions for non-Gaussianities. Note that already in standard DBI inflation with $\bar \Omega=0$ one has significant non-Gaussianities. Now due to explicit breakdown of four-dimensional diffeomorphism
one naturally expects to obtain non-Gaussianities with different shapes. The
situation here is to some extent similar to the EFTI literature on non-Gaussianities,
for example see \cite{Cheung:2007st, Senatore:2009gt, Bartolo:2010bj, Bartolo:2010di, Baumann:2011su}
and references therein. We would like to come back to the questions of cosmological perturbations and non-Gaussianities in a future publication.
\section{Summary and Discussions}
\label{summary}
In this paper a new model of DBI inflation is presented where the mobile brane is moving ultra relativistically inside a Lifshitz background with arbitrary scaling exponent $z$. This is a generalization of standard DBI inflation inside an AdS throat with $z=1$.
This work is mainly at the phenomenological level. As expressed previously, we do not have a rigorous construction of our Lifshitz throat within string theory set up. It is an open question as how one can embed the Lifshitz background rigorously in string theory. Furthermore, the concept of brane with the right RR coupling is not well-understood in this set up. We have employed the phenomenological approach that the Lifshitz throat with mobile branes, as required for brane inflation, can be embedded in principle in string theory. Furthermore, as in standard brane inflation, we assumed that
one can obtain the inflationary potential from the back-reactions of the mobile branes with the background fluxes and volume modulus. Whether or not the potential has the right form and the right couplings to support long enough period of inflation is the question of fine-tuning and on this issue our model is on the same footing as conventional models of brane inflation.
In this work we have not studied the slow-roll brane inflation. This is because the potential between brane and anti-brane required for slow-roll inflation, as in \cite{Kachru:2003sx}, is not understood for arbitrary value of $z$ in Lifshitz background. It would be interesting to calculate the inter-brane potential for generic value of $z$ so both slow-roll and fast-roll brane inflation can be combined for a more richer inflationary model building in this set up.
After dimensional reduction we found that the four-dimensional general covariance is broken explicitly both in the matter and the gravitational sectors. The reason is that we start from a theory where the Lorentz invariance is broken at the level of solution. Performing the dimensional reduction to four dimension, this manifests itself in parameter $\bar \Omega$ which measures the level of 4D diffeomorphism breaking. Depending on how Lifshitz throat is glued to the bulk of CY compactification, we find that $\bar \Omega$ can have either signs and can be order of few. The total action, including the general metric and matter perturbations, is obtained in Eq. (\ref{S-total}). One expects the model has interesting cosmological predictions, specially for the magnitude and the shapes of the non-Gaussianities. We would like to come back to these question in a future publication.
\section*{Acknowledgement}
We would like to thank P. Creminelli, R. Fareghbal, J. Maldacena,
A. Nicolis, L. Senatore and E. Silverstein for useful discussions. We also thank
M. M. Sheikh-Jabbari and the anonymous JCAP referee for the comments on perturbations in
the earlier version of the draft. We specially thank Kazuya Koyama for many insightful discussions and comments. H.F. would like to thank ICG for the hospitality during the revision of this work.
|
\section{\label{intro}Motivation}
Quantum state tomography \cite{book:stateestimation} is prone to errors,
of various origins.
For instance,
samples might be small,
the observables measured might not be informationally complete,
or
measurement devices might be inaccurate.
In this letter, I focus on one further source of systematic error:
The quantum system might be coupled in an uncontrollable fashion to a heat bath of fixed but unknown temperature (say, a piece of equipment or the surrounding air), whose effect may moreover
vary randomly from sample to sample
(say, because the duration of interaction varies between different runs of the experiment).
Transient contact with the bath then triggers partial relaxation (to varying extent) of the system towards thermal equilibrium.
In general, neither
the original quantum state of the system nor
the pertinent system-bath dynamics are known;
and so neither is
the target equilibrium state,
let alone the trajectory in state space leading from the original state towards the latter.
Under such adverse circumstances,
how can one hope to learn anything about the original quantum state?
In this letter,
I provide a simple, approximate scheme that allows one at least to constrain the unknown quantum state to a one-dimensional submanifold of state space,
based on the outcomes of multiple independent runs of the experiment.
The reconstruction of this one-dimensional constraint manifold assumes a generic relaxation dynamics and uses the maximum likelihood approximation.
\section{\label{relaxation}Generic relaxation dynamics}
In this section
I first discuss the generic relaxation dynamics close to equilibrium
and subsequently propose an extension to states further away.
Close to equilibrium, in the linear response regime,
states typically relax towards equilibrium exponentially,
\begin{equation}
\rho(t)-\sigma
=
\exp(-t/\tau_{\rm rel}) [\rho(0)-\sigma]
,
\end{equation}
on some pertinent time scale $\tau_{\rm rel}$.
Here $\sigma$ denotes the equilibrium state of the system.
In the setting considered here, both $\tau_{\rm rel}$ and $\sigma$ are fixed but unknown.
The relaxation time $\tau_{\rm rel}$ should be larger than the typical contact times with the bath so that relaxation of the system towards equilibrium is only partial.
Such exponential relaxation can be regarded as
a \textit{steepest-descent flow} of the relative entropy $S(\rho\|\sigma)$, in the following sense.
Given any set of observables $\{F_b\}$ which is informationally complete,
every state $\rho$ can be written in the Gibbs form
\begin{equation}
\rho = Z(\lambda)^{-1} \exp\left[ (\ln\sigma-\langle\ln\sigma\rangle_\sigma)-\sum_b \lambda^{b} F_b \right]
,
\label{gibbs}
\end{equation}
with properly adjusted Lagrange parameters $\{\lambda^{b}\}$
and partition function $Z(\lambda)$ \cite{ruskai:minrent,olivares+paris,PhysRevA.84.012101}.
(This is not to be confused with a thermal state:
A thermal state has the Gibbs form, too,
yet with only one observable (Hamiltonian) and Lagrange parameter (inverse temperature) in the exponent,
and with $\sigma$ replaced by the totally mixed state.)
The Lagrange parameters $\{\lambda^{b}\}$ and the expectation values $\{f_b:=\langle F_b\rangle_\rho\}$,
respectively,
constitute two possible choices for the coordinates in state space.
Associated with these coordinates are the respective basis vectors $\{\partial_b:=\partial/\partial\lambda^b\}$
and $\{\partial^b:=\partial/\partial f_b\}$,
as well as their one-form duals $\{d\lambda^a\}$ and $\{df_a\}$,
related via
$d\lambda^a(\partial_b)=df_b(\partial^a)=\delta^a_b$ \cite{schutz:book}.
The transformation from one coordinate basis to the other is effected by the correlation matrix
$C_{ab}:=\partial^2 \ln Z(\lambda)/\partial\lambda^a\partial\lambda^b$,
\begin{equation}
\partial_a= -C_{ab} \partial^b
\ ,\
\partial^a = -(C^{-1})^{ab} \partial_b
\end{equation}
(where I have adopted the convention that identical upper and lower indices are to be summed over),
and likewise the transformation of their duals,
\begin{equation}
df_a = -C_{ab} d\lambda^b
\ ,\
d\lambda^a = -(C^{-1})^{ab} df_b
.
\end{equation}
Being symmetric and positive, the correlation matrix also provides a natural Riemannian metric on state space,
the Bogoliubov-Kubo-Mori (BKM) metric \cite{bengtsson:book,balian:physrep,10.1063/1.530611,10.1063/1.531535,Jochen200083,grasselli:uniqueness}.
Its elements may be regarded as the components of the metric tensor
\begin{equation}
C:=C_{ab} d\lambda^a \otimes d\lambda^b
.
\end{equation}
The relative entropy $S(\rho\|\sigma)$ has the gradient \cite{PhysRevA.84.012101}
\begin{equation}
dS(\rho\|\sigma) = - \lambda^a df_a
,
\end{equation}
and so the pertinent steepest-descent curve (in the BKM metric introduced above) has a tangent vector
\begin{equation}
V \propto C^{-1}(-dS(\rho\|\sigma)) = -\lambda^a \partial_a
.
\end{equation}
Consequently, on a steepest-descent trajectory the Lagrange parameters must evolve according to
$d\lambda^b/dt \propto (-\lambda^a \partial_a)\lambda^b = -\lambda^b$
and thus relax exponentially towards $\lambda=0$, which corresponds to the state $\sigma$.
In the linear response regime,
exponential relaxation of the Lagrange parameters in turn implies exponential relaxation of expectation values,
and so indeed of the statistical operator,
Q.E.D.
Identifying thus the generic relaxation dynamics near equilibrium with a steepest-descent flow
opens the possibility to formulate a generic dynamics even in regions further away:
In the absence of detailed information about the system-bath dynamics,
and regardless of how far the state of the system might be from equilibrium,
I henceforth assume that
the generic effect of a bath on the system is to move its state by some (unknown) distance along the steepest-descent trajectory towards some (unknown) equilibrium state $\sigma$.
Such an extension of the steepest-descent paradigm to full state space has been studied before,
and in the special case of the classical Fokker-Planck equation (with a particular choice of metric)
has been shown
to reproduce the effective dynamics excactly
\cite{Jordan1997265,10.1137/S0036141096303359}.
In the setting considered here, the steepest-descent algorithm induces an effective quantum operation on the state space of the system.
Close to equilibrium,
this effective quantum operation is a completely positive map.
Further away from equilibrium, however,
it might no longer be linear and thus not completely positive;
which suggests that,
in contrast to a completely positive map,
the steepest-descent ansatz may presume non-negligible initial correlations between system and bath.
Such initial correlations are not surprising in view of the fact that the properties of the heat bath are not known \textit{a priori};
they simply reflect the fact
that learning something about the initial state of the system
will entail learning something about the state of the bath, too.
(There are also other reasons why one should expect the effective quantum dynamics to be nonlinear,
see e.g. Ref. \cite{PhysRevA.82.052119}.)
It remains to be seen whether for a given steepest-descent flow there exists always some compatible microscopic system-bath dynamics,
just as any completely positive map always arises from some underlying unitary evolution of the (initially uncorrelated) composite system-bath state.
I conjecture that this is the case, but defer its proof to future work.
Any steepest-descent trajectory towards $\sigma$ can be parametrized in the form
\begin{equation}
\rho(\gamma) \propto \exp\left[ (\ln\sigma-\langle\ln\sigma\rangle_\sigma)- \gamma G \right]
,
\label{trajectory}
\end{equation}
with some \textit{single} observable $G$ and associated Lagrange parameter $\gamma$.
I prove this assertion in two steps:
(i)
any curve parametrized in this form is a steepest-descent curve towards $\sigma$;
and
(ii)
for every initial state $\rho_0$ there exists a curve of this form passing through it.
The first statement becomes evident when one considers the above parametric form to be
a special case of the more general form (\ref{gibbs}),
with the pair $(\gamma,G)$ being one of the informationally complete $\{(\lambda^b,F_b)\}$ and
the Lagrange parameters associated with all $F_b\neq G$ equal to zero.
If these other Lagrange parameters are initially zero, they remain zero under the steepest-descent dynamics
$d\lambda^b/dt \propto -\lambda^b$.
So indeed, a curve parametrized as above describes a steepest descent towards $\sigma$.
As for the second statement,
for any initial $\rho_0$ (except for states on the boundary of state space)
the choice $G=\ln\sigma-\ln\rho_0$ will yield a steepest-descent curve passing through both
$\rho_0$ (at $\gamma=1$) and $\sigma$ ($\gamma=0$).
Given $\rho_0$ and $\sigma$,
this choice for $G$ is unique up to multiplicative and additive constants.
Given the imperfect data from multiple runs of the experiment that have been affected
(to varying degree) by uncontrolled contact of the system with the same unknown heat bath,
a precise reconstruction of the original quantum state is clearly not possible.
However, assuming that transient contact with the bath triggers a generic relaxation dynamics as described above,
it is at least possible to constrain the original quantum state to some steepest-descent trajectory.
Having established the parametric form (\ref{trajectory}) of such a trajectory,
the task is then to infer from the available experimental data the observable $G$
(up to multiplicative and additive constants)
as well as \textit{some} state (not necessarily the target state $\sigma$) which lies on this trajectory.
This is the inference task to which I turn in the next section.
\section{\label{theory}Inferring the constraint curve}
The experiment is run multiple times.
In the $i$th run, after the supposed disturbance by the bath,
one performs complete quantum state tomography on a sample of size $N_i$ and obtains the tomographic image $\mu_i$.
Different runs of the experiment are disturbed by the same bath but to varying extent (say, because the contact time varies between runs).
As a result,
the original quantum state is shifted on its relaxation trajectory towards equilibrium to varying degree;
and so rather than being clustered
around a single point (which would yield a unique state estimate),
the images $\{\mu_i\}$ are expected to be
spread out along this trajectory.
Assuming that the relaxation trajectory has the steepest-descent form (\ref{trajectory}),
the problem of inferring the observable $G$ from images that correspond to various values of $\gamma$
is analogous to the problem of estimating a Hamiltonian from thermal data at various temperatures.
Assuming that the tomographic images are not too far apart, and
expanding $G$ in terms of the informationally complete set $\{F_b\}$,
\begin{equation}
G=-\xi^b F_b
\end{equation}
(with the same index summation convention as above),
the log-likelihood of observing the data $\{\mu_i\}$ is asymptotically
(i.e., for sufficiently large sample sizes)
given by
\begin{equation}
L(\{\mu_i\}|\xi,\sigma)
\sim
(N/2)
[\langle\Gamma\rangle_{\xi} - (C^{-1})^{ab}\delta f_a(\xi,\sigma) \delta f_b(\xi,\sigma)]
,
\label{xi_likelihood}
\end{equation}
modulo additive terms that do not depend on $\xi$ or $\sigma$ \cite{PhysRevA.84.052101}.
Here
$N:=\sum_i N_i$,
and
\begin{equation}
\langle \Gamma\rangle_\xi
:=
\frac{\Gamma_{ab}\xi^a \xi^b}{C_{cd}\xi^c \xi^d}
\end{equation}
is the ``expectation value'' of the covariance matrix
\begin{equation}
\Gamma_{ab}:=\sum_{i} w_i (f_a^i-\bar{f}_a)(f_b^i-\bar{f}_b)
.
\end{equation}
The latter in turn depends on the deviations of the
sample means $f_b^i:=\langle F_b\rangle_{\mu_i}$
from their weighted averages $\bar{f}_b := \sum_i w_i f_b^i$,
with the $i$th sample carrying relative weight $w_i:=N_i/N$.
The correlation matrix $C_{ab}$ is evaluated at the center of mass
$\bar{\mu}:=\sum_i w_i \mu_i$
of the tomographic images.
Finally, the variations $\delta f$ are given by
\begin{equation}
\delta f_b(\xi,\sigma):=\langle F_b\rangle_{\bar{\pi}(\xi,\sigma)}-\bar{f}_b
,
\end{equation}
where $\bar{\pi}(\xi,\sigma)$ is the unique state of the form (\ref{trajectory})
which yields
$\langle G\rangle_{\bar{\pi}}=\langle G\rangle_{\bar{\mu}}$.
Unlike in the problem of estimating a Hamiltonian from thermal data where $\sigma$ corresponds to a given reference state,
the target state $\sigma$ is not fixed \textit{a priori}
but is itself a variable to be inferred.
To maximize the above log-likelihood as a function of both $\xi$ and $\sigma$,
the steepest-descent trajectory must maximize $\langle \Gamma\rangle_\xi$ as well as render $\delta f=0$.
The first requirement implies that
$\vec{\xi}$ must be the dominant eigenvector (i.e., the eigenvector associated with the largest eigenvalue) of $\bm{C}^{-1}\bm{\Gamma}$,
the product of the inverse correlation and covariance matrices.
Such an eigenvector condition is a familiar result
in principal component analysis \cite{roweis:em,roweis:unify,bishop:bayesian_pca,bishop:variational_pca,RSSB:RSSB196}.
The second requirement, on the other hand, is met by any $\sigma$ that ensures $\bar{\pi}=\bar{\mu}$,
and hence by any steepest-descent trajectory that passes through the center of mass $\bar{\mu}$.
Thus one can infer from the experimental data both a maximum likelihood estimate for $\vec{\xi}$, and hence for $G$,
and a state (namely $\bar{\mu}$) through which the steepest-descent trajectory must pass.
Together, these characterize the steepest-descent trajectory uniquely.
The original quantum state of the system must lie somewhere on this trajectory.
\section{\label{toy}Example: qubits}
The simplest example
is state tomography on an exchangeable sequence of qubits, emitted by some i.i.d. quantum source.
The tomographic experiment is performed several times,
with the $i$th run yielding the tomographic image $\mu_i$.
Rather than isotropically around a point, these images $\{\mu_i\}$ turn out to be clustered along a curve in state space;
say, along a straight line parallel to the $y$ axis of the Bloch sphere (Fig. \ref{qubit}).
This suggests that in between runs of the experiment,
there is some uncontrolled change in the environment which is characterized by a single parameter.
An obvious candidate is the coupling to an unknown heat bath,
with the contact time fluctuating in between runs.
For a qubit,
the informationally complete set of observables $\{F_b\}$ can be taken to be the Pauli matrices.
The associated parameter vector $\vec{\xi}$ points in the direction of the effective
``dissipative force'' which drives the qubit towards equilibrium.
Due to the nontrivial geometry of quantum states this force is in general \textit{not} parallel
to the spatial orientation of the data cluster
(here: the $y$ axis) as one might na{\"i}vely expect.
Rather, it is tilted against this axis by some angle $\phi$.
This angle can be calculated as follows.
Assuming that the measured data are perfectly aligned parallel to the $y$ axis as shown in Fig. \ref{qubit},
all entries of the covariance matrix vanish except for $\Gamma_{yy}\neq 0$.
And for the center-of-mass state $\bar{\mu}$ located in the $x-y$ plane as shown in Fig. \ref{qubit},
the inverse correlation matrix has the form
\begin{equation}
\bm{C}^{-1}(\bar{\mu})=
\left(
\begin{array}{ccc}
C^{-1}_{xx} & C^{-1}_{xy} & 0 \\
C^{-1}_{yx} & C^{-1}_{yy} & 0 \\
0 & 0 & C^{-1}_{zz}
\end{array}
\right)
.
\end{equation}
The requirement that
$\vec{\xi}$ be the dominant eigenvector of $\bm{C}^{-1}\bm{\Gamma}$ then implies
$\xi_z=0$ and $\xi_x/\xi_y=C^{-1}_{xy}/C^{-1}_{yy}$.
As a consequence, $\vec{\xi}$ is tilted against the $y$ axis by the angle
\begin{equation}
\phi=\arctan(C^{-1}_{xy}/C^{-1}_{yy})
.
\end{equation}
This tilting angle vanishes whenever the center of mass lies on one of the two ($x$ or $y$) axes.
Away from the axes, however, the tilting angle is non-zero, and increases as the center of mass moves closer to the surface of the Bloch sphere.
To illustrate the latter, I consider a center-of-mass state on the $x-y$ diagonal,
with azimuth $\pi/4$ and variable Bloch vector length $r$.
The relevant elements of the inverse correlation matrix are then
$C^{-1}_{yy}=1/(1-r^2)$ and $C^{-1}_{xy}=C^{-1}_{yy}-(\tanh^{-1}r)/r$.
This yields a function $\phi(r)$ which increases monotonically
from $\phi(0)=0$ to $\phi(1)=\pi/4$,
and which to a reasonable degree of accuracy
can be approximated by $\phi(r)\approx(\pi/4)r^2$.
Only near the origin of the Bloch sphere, therefore, and hence for highly mixed states,
does the inferred direction of the effective dissipative force coincide with the ``na{\"i}ve'' estimate based on the spatial orientation of the data cluster.
For states that are (nearly) pure,
on the other hand,
the reconstruction scheme presented here may yield a direction $\vec{\xi}$ which differs significantly from that na{\"i}ve estimate.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=9cm]{qubit}
\end{center}
\caption{First quadrant of a two-dimensional section ($z=0$) of the Bloch sphere.
The black dots indicate the tomographic images $\{\mu_i\}$ obtained in different runs of the experiment, and the small circle their center of mass $\bar{\mu}$.
The Bloch vector associated with the center of mass is assumed to have no $z$ component ($\langle\sigma_z\rangle_{\bar{\mu}}=0$),
azimuth $\pi/4$, and length $r$.
Even though the images are aligned parallel to the $y$ axis,
the inferred direction $\vec{\xi}$ of the effective force which drives the relaxation towards equilibrium is tilted against that axis by an angle $\phi$.
This angle grows from $\phi=0$ at $r=0$ to $\phi=\pi/4$ at $r=1$.
}
\label{qubit}
\end{figure}
\section{\label{discussion}Conclusions}
Whenever an i.i.d. sequence of quantum systems is subjected to disturbance by the same heat bath,
yet to an extent that varies randomly between samples,
multiple runs of quantum state tomography are expected to yield data spread out along a curve in state space.
From the shape and location of this sprawl,
one can infer the system's effective relaxation dynamics
under the influence of the bath.
Owing to the nontrivial geometry of quantum states,
the result of this inference can be rather counterintuitive.
In the qubit example, the inferred direction of the effective dissipative force generally deviated from the principal axis of the data cluster.
There are several ways to extend the results of the present letter,
which are left to future work.
(i)
So far the reconstruction scheme for the effective relaxation dynamics does not take into account the prior distribution $\mbox{prob}(\xi,\sigma)$ of its parameters.
Doing so will lead to a Bayesian modification of the maximum likelihood framework presented here.
A full Bayesian analysis should also include a study of the error bars on $\xi$ and $\sigma$.
(ii)
Many physical systems, especially larger ones,
exhibit not a single relaxation time but a whole hierarchy of time scales pertaining to the relaxation of different sets of degrees of freedom.
In this case thermalization occurs in stages,
on successively longer time scales \cite{Balian1987229,rau:physrep}.
Provided the typical contact times with the bath are shorter than the smallest scale in that hierarchy,
the reconstruction scheme presented here is still valid.
It pertains then to the first thermalization stage,
with $\sigma$ being no longer the equilibrium state but the target state of this first stage.
If contact times vary widely across different relaxation time scales, however,
the scheme must be adapted.
(iii)
The system might be disturbed by not just one but several baths, each to independently varying degree,
or the impact of a single bath might be governed by more than one parameter.
Then, rather than along a curve, tomographic images will be spread out in some higher-dimensional submanifold of state space.
Inferring the dimensionality of this submanifold, as well as the associated parameters,
will require a further generalisation of the scheme presented here.
(iv)
Reconstructing the relaxation trajectory (or submanifold) is the first step towards the reconstruction of the system's {unperturbed} quantum state.
I argued that this unperturbed quantum state must lie somewhere on the relaxation trajectory (or submanifold);
but its precise location will depend on further assumptions, not considered in the present letter,
about the distribution of the contact times and possibly other parameters of the bath.
(v)
Finally,
I consider it worthwhile to study in more detail the conceptual underpinning as well as possible modifications of the steepest-descent paradigm.
And on a speculative note,
taking the steepest-descent paradigm at face value and identifying (some suitable function of) relative entropy as actual ``time,''
one might even be tempted to try to establish a link to (equally speculative) ideas in other areas of physics that aim to reduce the notion of time to a distance between configurations \cite{barbour:timeless} or thermal properties \cite{0264-9381-11-12-007}.
\bibliographystyle{elsarticle-num}
|
\section{Introduction}
\label{intro}
Scaling properties of the local surface roughness $w$ and of the height-height correlation
function (HHCF) $C$ are very useful to understand the growth dynamics of thin films and
other deposits \cite{barabasi,krug,krim}. The usual approach is to measure exponents
from plots of $w$ or $C$ as a function of the box size $r$ (roughness exponent) or time $t$
(growth exponent) and to relate their values
to some universality class of growth \cite{barabasi}.
However, a very small number of systems exibit simple scaling features to match those
theories. For instance, the presence of grains in the film surface leads to a crossover
between two regimes where $w$ increases with $r$ with different roughness exponents,
$\alpha_1$ and $\alpha_2$, as illustrated in Fig. 1
\cite{lita,kleinke1999,ebothe2003,vazquez,mendez,otsuka,vasco}.
For the HHCF, the same crossover occurs with exponents $\chi_1$ and $\chi_2$..
Similar crossover is observed in other systems, such as fresh snow on the ground and
pyroclastic deposits on volcanic surfaces \cite{manes,mazzarini}.
In Ref.\protect\cite{grains}, the crossover with $\alpha_1\approx 1$
was shown to be a geometric effect of the grainy surface structure and of the gliding box method
(analogous result is obtained with the box counting method).
It was also shown that the crossover took place when $r$ was close to the average grain size.
If the grain surface is flat, $\alpha_1$ is very close to $1$, while for rounded grains
it decreases to values close to $0.85$ \cite{grains}. These results match
those of a large number of experimental works
\cite{lita,kleinke1999,ebothe2003,vazquez,mendez,otsuka,vasco}.
However, other experimental works show film surfaces with grainy structure, the same
crossover in roughness scaling or HHCF, but with much smaller exponents $\alpha_1$
\cite{vazquezSS,tersio,mendomeJCG,mendomeMCP,hiane,nara2005,nara2007,naraNano2007,marta2002}.
The usual interpretation for those exponents is that small
scale surface features are determined by a different growth dynamics.
Indeed, even the crossover with $\alpha_1\approx 1$ was already interpreted as an
anomalous scaling, with $\alpha_1$ being called local roughness exponent (denoted
$\alpha_{loc}$) and $\alpha_2$ called global roughness exponent. For these reasons,
in many systems it is still unclear whether a crossover similar to that in Fig. 1
should be interpreted as a purely geometric effect or as a consequence of a
competitive growth dynamics.
\begin{figure}[!b]
\includegraphics[width=7cm]{Fig1.eps}
\caption{Typical behavior of the local roughness as a function of box size in grainy surfaces.}
\label{fig1}
\end{figure}
Here we study several growth models with grainy surface features to show the
possible effects of the grain shape, of the method of calculation of
averages of squared quantities, of the working quantity ($w$ or $C$) and
of the universality class of the growth process.
For all growth models, grain shapes, and methods of analysis, we observe
crossovers at box sizes very close to the average grain size.
We also show that a very broad range of $\alpha_1$ can be found, depending on the
grain shape and the working quantity, but independently of the universality class of growth,
which determines only the value of $\alpha_2$. Similar conclusions are obtained for the
exponents $\chi_1$ and $\chi_2$. The comparison with experimental works with
several materials and deposition methods gives additional support to the geometric
interpretation of the crossover in those systems.
The rest of this work is organized as follows. In Sec. \ref{model} we define average quantities
and present the growth models. In Sec. \ref{theoretical}, we recall the results of some
exactly solvable models with grains at the surface, which explain the crossover with
$\alpha_1\approx 1$ and $\chi_1\approx 0.5$ (with the usual definition of the HHCF).
In Sec. \ref{shape} we analyze the effects
of the grain shape, particularly some very sharp grains, considering models in
different universality classes. In Sec. \ref{experimental}, we show the applications
of our approach to real films. In Sec. \ref{conclusion}, we summarize our
results and present our conclusions.
\section{Definition of average quantities and models}
\label{model}
First we define the average quantities analyzed in this work.
The surface roughness in square boxes of size $r$ at time $t$ is usually defined as
\begin{equation}
w{\left( r,t\right)}\equiv {\left<
{\overline{{\left( h-\overline{h}\right)}^2}}^{1/2}\right>} .
\label{defw}
\end{equation}
The overbars in Eq. (\ref{defw}) denote averages of the height $h$ inside a given
box position (spatial average)
and the angular brackets represent the configurational average as the box
scans the whole surface of a deposit. This is called gliding box method, in which the scanning
box moves one pixel each time it performs a new spatial average.
In box counting methods, the surface is divided in nonintersecting boxes for the
configurational average.
Alternatively, some authors define the roughness as
\begin{equation}
w' \equiv {\left< \overline{{ \left( h -\overline{h}\right) }^2} \right>}^{1/2} ,
\label{defw1}
\end{equation}
i. e. they calculate the configurational average of the square height fluctuation and
take the square root of that average.
When several images of a deposit are available, or several configurations are grown with the
same model, these different samples also contribute to the above configurational averages.
For window sizes below the grain size, the roughness scales as
\begin{equation}
w{\left( r,t\right)} \sim r^{\alpha_1} ,
\label{defalpha1}
\end{equation}
which defines the initial roughness exponent $\alpha_1$ (Fig. 1).
The height-height correlation function (HHCF) at distance $r$ and time $t$ is usually defined as
\begin{equation}
C\left( r,t\right) \equiv {\langle {\left[ h\left( r_0+r,t\right) -h\left( r_0,t\right)
\right] }^2\rangle}^{1/2} ,
\label{defcorr}
\end{equation}
with configurational averages taken over all different initial positions $r_0$.
Alternatively, it can be defined as
\begin{equation}
C'\left( r,t\right) \equiv {\langle {| h\left( r_0+r,t\right) -h\left( r_0,t\right) |}
\rangle} ,
\label{defcorr1}
\end{equation}
which corresponds to an interchange of the
configurational average and the calculation of the square root
in Eq. (\ref{defcorr}). In this sense, the calculation of $C(r,t)$
parallels that of $w'$, while the calculation of $C'(r,t)$ parallels that of $w$.
For window sizes below the grain size, the HHCF scales as
\begin{equation}
C{\left( r,t\right)} \sim r^{\chi_1} ,
\label{defchi1}
\end{equation}
which defines the initial roughness exponent $\chi_1$ for that function.
For window sizes much larger than the grain size (i. e. $r\gg r_c$ - see Fig. 1),
a surface obeying normal scaling has
$w\sim r^{\alpha_2}$ and $C\sim r^{\chi_2}$, with $\alpha_2=\chi_2$.
The quantities $w'$ and $C'$ obey the same scaling. Those exponents are
representative of the large lengthscale kinetics governing the growth process. Typical
examples of growth kinetics are those of Edwards-Wilkinson (EW)
\cite{ew}, of Kardar-Parisi-Zhang (KPZ) \cite{kpz},
and the diffusion-dominated ones, linear (Mullins-Herring - MH) \cite{mh} or
nonlinear (Villain-Lai-Das Sarma - VLDS) \cite{villain,laidassarma}.
Now we present the models for growth of thin films with grains at the surface.
Intrinsic corrections to scaling for large $r$ and large $t$ should be avoided in those
models, so that any crossover is solely due to the grainy structure.
This request excludes the grain deposition models
introduced in Ref. \protect\cite{grains} and related ballistic-like models
\cite{bbd,intrinsic} because they have remarkable scaling corrections.
On the other hand, some models with smooth surfaces and particle enlargement presented in
Refs. \protect\cite{vazquez,grains} satisfy that condition. They are described below.
The first model has KPZ kinetics. The first step is to grow a deposit with cubic particles
of unit size
following the rules of the restricted solid-on-solid (RSOS) model: the aggregation
of the incident particle is accepted only if the height differences of nearest neighbors
are always $0$ or $1$ (otherwise the aggregation attempt is rejected) \cite{kk}.
We recall that $\alpha_2=\chi_2\approx 0.39$ for the KPZ class in
two-dimensional substrates \cite{eqkpz}.
The second model has VLDS kinetics. The initial deposit is grown with the rules of the
conserved RSOS model, where the incident particle executes a random walk between
neighboring columns until finding a column where it can aggregate respecting the conditions
on height differences \cite{crsoskim,crsos}.
We recall that $\alpha_2=\chi_2\approx 0.67$ for the VLDS
class in two-dimensional substrates \cite{janssen,crsos}.
After growing the initial deposit, with KPZ or VLDS model,
the size of each particle is enlarged by a factor $l$, i. e. each particle is
transformed in a cubic grain of side $l$. Most of our simulations are performed with
$l=32$.
The final step is replacing the top cub grains (surface grains) by rounded or sharp
structures. Three shapes are used:
semi-ellipsoids of horizontal radius $l\sqrt{2}/2$ and vertical radius $h$,
cones with that radius and height $h$, and pyramids of square basis of side $l$
and height $h$. They are illustrated in Fig. \ref{fig2}. Several values of $h$ are considered
for each shape, typically between $l$ and $3l$.
\begin{figure}[!h]
\includegraphics[width=9cm]{Fig2.eps}
\caption{Shapes of surface grains after enlargement of the original deposits:
a) semi-elliptical, b) conical, c) pyramidal.
Semi-ellipsoids and cones are cut at the sides so that their basis are squares
of side $l$ that fit the shape of the cubic grain at their bottom.
This is the reason for radius $l\sqrt{2}/2$ of their basis.}
\label{fig2}
\end{figure}
In the scaling of $w$ or $C$, the role of the height $h$
is measured relatively to the height of the surface steps, which is $l$. The
horizontal scaling factor is also $l$ for the cubic grains, but this is not important
for the scaling exponents. For instance, if the grains were constructed with the shape
of paralelepids of height $l$ and horizontal sides $l_{\|}$,
the scaling exponents would not change.
Thus, the aspect ratio of the grains considered here is not a limitation of the model.
The simulations of the KPZ and VLDS models
were performed in square substrates (three-dimensional
deposits) of lateral size $L=128$ at times of order ${10}^4$. For the RSOS model, it
corresponds to approximately $5\times {10}^3$ layers of unit size particles; for the
CRSOS model, corresponds to ${10}^4$ layers. After replacement of the original particles
by grains of size $l=32$, the deposits have lateral size $4096$.
Simulations in smaller sizes ($L=64$ and $L=32$ for the original models)
and different grain size ($l=16$) give similar results for all exponents,
indicating that finite-size and finite-time effects are negligible.
\section{Theoretical predictions for $\alpha_1$ and $\chi_1$}
\label{theoretical}
As the scanning box glides along the surface, it frequently encloses high surface
steps created between neighboring grains. These are the box positions where the largest
height fluctuations are encountered, thus they give the main contribution to the roughness
(Eqs. \ref{defw} or \ref{defw1}).
If the box has size $r$ (i. e. $r$ pixels in
each direction), then the number of box positions that involve each high step is
proportional to $r$.
Thus, the configurational average of Eq. (\ref{defw}) gives roughness $w$ proportional
to $r$. This gives $\alpha_1 =1$, as explained in
Ref. \protect\cite{grains} and confirmed by simulations of several models.
When Eq. (\ref{defw1}) is used, ${w'}^2$ is a configurational average.
The main contribution to that average also comes from box positions
enclosing high surface steps, thus, that average is proportional to $r$.
This gives $w'$ proportional to $r^{1/2}$, i. e., $\alpha_1=1/2$.
These results are confirmed by our simulations of
the RSOS model with cubic grains, as shown in Fig. 3a. It clearly shows the
remarkable difference in the scaling of $w$ and $w'$ for box sizes smaller than
the grain size, while the same exponent $\alpha_2$ after the crossover
represents the universality class of the process.
\begin{figure}[!t]
\includegraphics[width=7cm]{Fig3.eps}
\caption{Scaling with the window size $r$ of data for the KPZ model with cubic grains:
a) $w$ (blue squares) and $w'$ (red triangles); and b) $C'$ (blue squares) and $C$ (red triangles).}
\label{fig3}
\end{figure}
Similar situation is observed with the HHCF.
Again the main contribution for the configurational average comes from box positions
which involve high surface steps, thus this average is proportional to $r$. With
the most used definition of that function (Eq. \ref{defcorr}),
we have $C(r,t)$ proportional to $r^{1/2}$,
thus the crossover takes place with $\chi_1 =1/2$. Instead, if the scaling of
$C'(r,t)$ is analyzed, we expect $\chi_1=1$.
Simulations of the RSOS model with cubic grains show the predicted crossover,
as illustrated in Fig. 3b.
The exponent $\chi_1$ is very close to $1/2$ for $C(r,t)$
and slightly below $1$ for $C'(r,t)$. Again, the universal exponent
$\chi_2$ is obtained after the crossover; as expected, $\alpha_2\approx \chi_2$.
With the usual definitions of surface roughness ($w$ - Eq. \ref{defw}) and HHCF
($C$ - Eq. \ref{defcorr}), the roughness exponents measured
before the crossover ($\alpha_1$, $\chi_1$) are different. This contrasts with
the expected universality after the crossover ($\alpha_2\approx \chi_2$).
Our analysis show that those discrepancies are effects of
the grainy morphology and the calculation method, in particular the order of
calculation of square root and configurational average in Eqs. (\ref{defw})
and (\ref{defcorr}).
\section{Effects of grain shape}
\label{shape}
Rounding of the surface grains may lead to $\alpha_1$ between $0.85$ and $1$,
as shown in Ref. \protect\cite{grains}.
However, the replacement of the cubic grains by the rounded or sharp structures in Fig. 2,
with $h\geq l$, leads to much more drastic changes in the initial exponents of $w(r,t)$ and
$C(r,t)$.
This result is illustrated in Figs. 4a and 4b for films grown with the KPZ model
and pyramidal grains of height $h=64$: $\alpha_1$ decreases to $0.71$ and $\chi_1$
increases to $0.61$ (while $\alpha_2\approx 0.39$). Fig. 4b also shows the
formation of a plateau in the HHCF before the second scaling regime, which is
characteristic of all sharp grains with large heights.
\begin{figure}[!b]
\includegraphics[width=7cm]{Fig4.eps}
\caption{a) Roughness ($w$) and b) HHCF ($C$) as a function of the window size $r$, for the KPZ
model with pyramidal grains of heigth $h=64$.}
\label{fig4}
\end{figure}
In Table I, we show the values of $\alpha_1$ and $\chi_1$ obtained for semi-elliptical,
conical and pyramidal grains with several heights.
A remarkable result is that films grown with the VLDS model have the same
exponents $\alpha_1$, $\chi_1$ up to the second decimal place, despite the
significant change in the asymptotic roughness exponent ($\alpha_2\approx 0.67$).
In Figs. 5a and 5b, we show results for conic grains
with height $h=32$, which give $\alpha_1= 0.809$ and $\chi_1=0.539$.
Those values are close to the KPZ values shown in Table I for the same grains.
The main differences from the models with KPZ scaling are that the
change in the slope of the roughness plot is smaller and there is a slope increase
in the HHCF plot when passing from the first to the second scaling regime.
\begin{figure}[!t]
\includegraphics[width=7cm]{Fig5.eps}
\caption{a) Roughness ($w$) and b) HHCF ($C$) as a function of the window size $r$, for the VLDS
model with conical grains of heigth $h=32$.}
\label{fig5}
\end{figure}
Table I shows that $\alpha_1$ is much smaller than the limit $0.85$ obtained in
previous work \cite{grains} for many grain shapes, particularly for sharp conic and pyramidal
grains. With the structures studied here, the lower limit is
close to $0.71$, obtained with pyramidal grains. For $h\leq 3l$, the general trend is that
the increase of $h$ leads to decrease of $\alpha_1$. For larger $h$ (not shown in Table I),
a very slow increase of $\alpha_1$ towards $1$ is observed.
\begin{table}[!t]
\begin{center}
\begin{tabular}{cccccccc}
\hline\hline
$h$ & & $32$ & & $64$ & & $96$ & \\
\hline
$\alpha_{1}^{SE}$ & & $0.826$ & & $0.773$ & & $0.763$ \\
$\chi_{1}^{SE}$ & & $0.504$ & & $0.551$ & & $0.589$ \\
\hline
$\alpha_{1}^C$ & & $0.806$ & & $0.768$ & & $0.768$ \\
$\chi_{1}^C$ & & $0.539$ & & $0.633$ & & $0.694$ \\
\hline
$\alpha_{1}^P$ & & $0.755$ & & $0.710$ & & $0.708$ \\
$\chi_{1}^P$ & & $0.535$ & & $0.606$ & & $0.645$ \\
\hline\hline
\end{tabular}
\caption{Exponents obtained from $w$ ($\alpha_1$) and $C$ ($\chi_1$) in KPZ films with
semi-elliptical (SE), conical (C) and piramidal (P) grains.}
\label{table1}
\end{center}
\end{table}
The relative changes in $\chi_1$ are much larger, attaining almost $40 \%$
for conic grains with $h=3l$ (see Table I). Indeed, this is the grain shape that provides higher
deviations from the flat grain value $\chi_1 =0.5$. A monotonic increase of $\chi_1$
is observed when taller grains are studied.
The above results show that sharp grain shapes bring closer the exponents
$\alpha_1$ and $\chi_1$, in contrast with the very different values for flat grains
($1$ and $0.5$, respectively - Sec. \ref{theoretical}).
In some cases, they are surprisingly close; for instance,
they differ only $10\%$ for conic grains with $h=3l$.
In Table II, we show exponents $\alpha_1$ and $\chi_1$ obtained from
the scaling of $w'(r,t)$ and $C'(r,t)$. They should be compared
with the respective flat grain values $0.5$ and $1$.
Comparison of results in Tables I and II show that sharp grain shapes also bring
closer the values of $\alpha_1$ measured from $w$ and $w'$ scaling, which are
very different for flat grains ($1$ and $0.5$, respectively - Sec. \ref{theoretical}).
It is particularly interesting to observe that $\alpha_1$ differs only $3\%$
when calculated from $w$ or $w'$ in films with pyramidal shapes with $h=3l$.
These values may be incorrectly interpreted as true roughness exponents because the
same $\alpha_2$ is expected for $w$ and $w'$. This type of erroneous interpretation
can be avoided if one accounts for the effects of a wide range of
grain shapes and sizes and investigates other quantities, such as HHCF.
\begin{figure}[!b]
\includegraphics[width=8cm]{Fig6.eps}
\caption{Film surface with $1/4$ of the grains flat and $3/4$ pyramidal with heights $h=32$,
$h=64$, and $h=96$ equally distributed.}
\label{fig6}
\end{figure}
The crossover size $r_c$ is defined at the intersection of the linear fits of the initial regime
and the second scaling regime of roughness or HHCF, as illustrated in Fig. 1..
Despite the wide range of values of $\alpha_1$ and $\chi_1$ shown in Tables I and II,
a remarkable result is that $r_c$ is always very close
to the grain size $l$, for KPZ and VLDS models.
Using $l=32$, our estimates range between $r_c=30$ and $r_c=34$,
which corresponds to a maximum difference of $7\%$. Consequently, $r_c$ can always be used as a
reliable estimate of the grain size.
In the above models, we considered surfaces with uniform grain height. However, we
also analyzed the effect of distributions of grain heights, since this is the situation
in real surfaces. In all cases,
we observe that the exponents $\alpha_1$ and $\chi_1$ are near the averages of those
obtained with a single value of grain height.
\begin{table}[!t]
\begin{center}
\begin{tabular}{cccccccc}
\hline\hline
$h$ & & $32$ & & $64$ & & $96$ & \\
\hline
$\alpha_{1}^{SE}$ & & $0.523$ & & $0.576$ & & $0.621$ \\
$\chi_{1}^{SE}$ & & $0.754$ & & $0.711$ & & $0.705$ \\
\hline
$\alpha_{1}^C$ & & $0.554$ & & $0.650$ & & $0.719$ \\
$\chi_{1}^C$ & & $0.712$ & & $0.693$ & & $0.710$ \\
\hline
$\alpha_{1}^P$ & & $0.556$ & & $0.638$ & & $0.687$ \\
$\chi_{1}^P$ & & $0.664$ & & $0.633$ & & $0.640$ \\
\hline\hline
\end{tabular}
\caption{Exponents obtained from $w'$ ($\alpha_1$) and $C'$ ($\chi_1$) in KPZ films with
semi-elliptical (SE), conical (C) and piramidal (P) grains.}
\label{table2}
\end{center}
\end{table}
An example of a film surface with such random grain distribution is shown in Fig. 6:
$1/4$ of the grains are flat and $3/4$ have pyramidal shape, with equally distributed heights $h=32$,
$h=64$, and $h=96$. For that surface, we obtain $\alpha_1=0.742$ and $\chi_1=0.554$,
which is close to the average of the results in Table I for those shapes.
\section{Comparison with experimental results}
\label{experimental}
In the experimental works discussed below, the exponent $\alpha_1$ defined here is
frequently named local roughness exponent $\alpha_{loc}$, as a reference to the small
lengthscale behavior and/or to a possible anomalous scaling.
Several experimental works have already shown the crossover of Fig. 1 with $\alpha \approx 1$,
which is explained by the growth models with flat or slightly rounded grains \cite{grains}.
Among those works, we highlight the study of rf sputtered
$LiCoO_x$ films by Kleinke et al \cite{kleinke1999}, which gives $0.91\leq \alpha_1\leq 0.95$;
the spray pyrolysis growth of $ZnO$ films by Eboth\'e et al \cite{ebothe2003}, which gives
$0.94\leq \alpha_1\leq 0.97$ for high flow rates; the electrodeposition of cooper by
Mendez et al \cite{mendez} and of gold by V\'azquez et al \cite{vazquez},
which give $\alpha_1=0.87\pm 0.06$
and $\alpha_1=0.90\pm 0.06$, respectively; the electrochemical
roughening of silver electrodes by Otsuka and Iwasaki \cite{otsuka}, which gives
$\alpha_1$ between $0.95$ and $0.98$; and the pulsed laser deposition of
$La$ modified-$PbTiO_3$ films of Vasco et al \cite{vasco}, which have $\alpha_1 = 1$.
However, many works show the same crossover with exponents $\alpha_1$
between $0.7$ and $0.85$, and surface images confirm the presence of grains
of approximately conic or pyramidal shape, much higher than the
steps between neighboring grains. These features are observed in films
of various materials and substrates, deposited with different techniques. This justifies
our approach with geometrical models, independently of the particular growth
dynamics.
Among the applications to inorganic materials, we find some vapor deposited gold films
by Vazquez et al, which have $\alpha_1\approx 0.83$ - see Fig. 1c and Fig. 3 of Ref.
\protect\cite{vazquezSS}. One of the niquel oxide film samples deposited by sputtering
in Ref. \protect\cite{tersio} have $\alpha_1 =0.70$, and the AFM image show the
qualitative features of our models with sharp grainy structure. Nearly the same
exponent ($\alpha_1=0.71$) is obtained with $Ni$ films electrodeposited on
indium tin oxide substrates in Refs. \protect\cite{mendomeJCG,mendomeMCP}.
Several $Ni-Zn$ alloy films of
Ref. \protect\cite{hiane} show the crossover in roughness scaling, with most estimates
of $\alpha_1$ in the range $[0.80,0.83]$. This is consistent with
our models of semi-elliptical grains of lower $h$, and the images actually show a
smooth grain morphology.
The same features are also observed in organic materials. Films formed
with bilayers of poly(allylamine hydrochloride) and a side-chain-substituted azobenzene
copolymer (Ma-co-DR13), after deposition of $10$ or $20$ bilayers, show grains with a
broad size distribution, and the
initial roughness exponents $0.81$ and $0.79$ \cite{nara2005}.
AFM images of chemically deposited polyaniline thin films on glass substrates \cite{nara2007}
have similar features, but, as far as we know, roughness scaling was not studied
with those images. The surface of Langmuir–Blodgett films of polyaniline and a
neutral biphosphinic ruthenium complex (Rupy) of Ref. \protect\cite{naraNano2007}
also show those grainy features with
some high peaks, and initial roughness exponents are in the range
$0.66\leq \alpha_1\leq 0.81$ for thicknesses between $1$ and $21$ layers. However,
most estimates of $\alpha_1$ are between $0.72$ and $0.76$ \cite{naraNano2007},
in good agreement with our results for pyramidal grains.
It is interesting to observe that
some surface images shown in Refs. \protect\cite{mendomeJCG,mendomeMCP,nara2005,nara2007,naraNano2007}
have features similar to the model illustration in Fig. 6 (in most cases without the flat grains).
This comparison reinforces our interpretation of the exponents measured in those works.
Similar results are obtained in etching of silicon surfaces
in Ref. \protect\cite{marta2002}: $\alpha_1$ is found between $0.70$ and $0.87$ when the
$(111)$ surface is etched by an $NaOH$ solution in contact with a non-saturated aqueous
environment.
It is also important to recall that there are systems with crossover in the roughness
scaling which do not show the sharp grainy features of our models, and consequently
deserve separate investigation. For instance, the images of another sample from
Ref. \protect\cite{tersio} does not show those features, but the roughness shows a
slow crossover with $\alpha_1=0.52$. Pyroclastic deposits of Mt. Etna show
roughness scaling crossover with $\alpha_1$ between $0.47$ and $0.67$,
but the images do not support modeling by grainy structures \cite{mazzarini}..
There are also systems with sharp grainy structures and small $\alpha_1$, such
as some $Ni$ films of Ref. \protect\cite{mendomeMCP} ($0.55\leq\alpha_1\leq 0.61$),
which also would deserve a separate investigation (those films have
$0.12\leq\alpha_2\leq 0.22$, which also cannot be easily explained with the well known
kinetic growth theories \cite{barabasi}).
The crossover in HHCF scaling obtained in some systems can also be related to our models.
For instance, Manes et al \cite{manes} used HHCF as a measure of fresh snow roughness and
obtained $\chi_1$ between $0.58$ and $0.62$ in a set of five experiments. These values
are consistent with our model with very high semi-ellipsoidal grains ($h=3l$) or with
conic or pyramidal grains with $h=2l$ or less.
Again, there are also systems where a crossover of HHCF scaling is observed but whose
images do not show the features of our models. An example is Ref. \protect\cite{martasergio},
where $\chi_1 =0.84$ was obtained for paraphin films deposited on stainless steel covered
with amorphous carbon.
Results of the recent work on pentacene island growth on stepped oxide surfaces \cite{conrad} can also
be related to our models. First, for long lengths, the HHCF has exponent $2\chi =1$, which
is expected for height fluctuations dominated by the surface steps; indeed, arguments analogous to those
for flat grains (Sec. \ref{theoretical}) give $\chi =1/2$. However, for small lengths, height
fluctuations in the surface terraces (due to pentacene islands) lead to the increase of
the HHCF exponent to the range $[0.69,0.8]$. Recent works showing evidence of anomalous scaling
in organic and inorganic film surfaces
also give estimates of HHCF exponents above $1/2$ at short lengthscales \cite{kim,zhang}. Although
both short and long range dynamics may be much more complex than in our models, a simple geometric
interpretation of the short range exponents may also be considered due to the presence of grainy
structures in the surface images.
\section{Conclusion}
\label{conclusion}
We extended the work on growth models with grainy surfaces to analyze the
effects of the grain shape, of the method of calculation of
averages of squared quantities, of the working quantity (roughness or HHCF) and
of the universality class of the growth process.
For all models, grain shapes, and methods of analysis, we observe
crossovers at box sizes very close to the average grain size.
We also show that a very broad range of the initial exponent $\alpha_1$ is found
for the roughness scaling, decreasing from $1$ for flat grains to $0.71$ for some sharp pyramidal
grains. The initial exponent $\chi_1$ of HHCF scaling increases from approximately
$0.5$ for flat grains to values larger than $0.7$ for sharp conic grains.
Simulations of KPZ and VLDS models show that the universality class has no significant
effect on the estimates of $\alpha_1$ and $\chi_1$.
The range of $\alpha_1$ presented here explains results of some recent experimental
works with different materials and deposition methods. This gives additional support
to the geometric interpretation of the crossover in roughness scaling for a
variety of systems.
\acknowledgments
FDAAR acknowledges support from CNPq and Faperj (Brazilian agencies).
|
\section{Introduction}
Multiferroic materials, having coexisting magnetism and ferroelectricity,
are of great technological and fundamental
importance,~\cite{Cheong,Ramesh,Khomskii1} given the prospect
of controlling charges by applying magnetic fields and spins by voltages.
BiCoO$_3$ was recently synthesized by a high-pressure (HP)
technique,~\cite{Belik} and it has been suggested to be a promising multiferroic
material by Uratani $et$ $al.$~\cite{Uratani} and by Ravindran
$et$ $al.$~\cite{Ravindran} both through first-principles Berry-phase calculations.
BiCoO$_3$ has a giant tetragonal
lattice distortion of $c$/$a$ = 1.27 with remarkable off-center atomic
displacements (see the inset of Fig. 1), and it is an insulator having C-type
antiferromagnetism below
470 K---the antiferromagnetic (AF) $ab$ layers stacking ferromagnetically along
the $c$ axis.~\cite{Belik}
It was proposed~\cite{Uratani,Oka,Okuno} that the giant tetragonal
distortion originates from lifting of the orbital degeneracy of the high-spin
(HS, $S$=2) Co$^{3+}$ ions and is stabilized by the subsequent $xy$-type
ferro-orbital ordering. Note that orbitally degenerate transition-metal
oxides quite often display an orbital ordering (OO) but ferroelectric (FE) materials
out of them are rare, as ferroelectricity and magnetism seem,
and actually in most cases,
to exclude each other.~\cite{Hill,Khomskii2}
Therefore, the proposed mechanism for the giant FE distortion appears
not straightforward.
Using fixed-spin-moment density-functional calculations,
Ravindran $et$ $al.$~\cite{Ravindran} predicted that there is a giant magnetoelectric
coupling in BiCoO$_3$: an external electric field (or a small
volume compression of $\sim$5\%) can induce a strong magnetic
response by changing the magnetic Co$^{3+}$-HS state in the
FE phase into a nonmagnetic
low-spin (LS, $S$=0) state in a paraelectric (PE) phase.
A corresponding HS-insulator/LS-metal
transition was also suggested.~\cite{Ravindran,Ming} In sharp contrast, a very recent
HP study~\cite{Oka} showed
that BiCoO$_3$ even under 6 GPa with a large volume decrease of 18\% is still semiconducting.
Note however that controversial spin states, both LS and intermediate-spin (IS, $S$=1),
were suggested for the HP phase.~\cite{Oka}
In the present work, we seek the origin of the giant tetragonal FE
distortion in the ambient phase
of BiCoO$_3$ and identify the nature of the pressure induced
spin-state transition, using two sets of configuration-state-constrained
GGA+$U$ (generalized gradient approximation plus Hubbard $U$) calculations.
Our results show that the giant tetragonal distortion is driven by
a strong Bi-O covalency (rather than by the aforementioned
lifting of orbital degeneracy) and is further stabilized by an $xy$-type OO of the
HS Co$^{3+}$ ions.
Moreover, we find that the pressure-induced spin-state transition is via a
mixed HS+LS state, which accounts for the available experimental
results consistently and disproves a recent prediction of a readily switchable
HS-LS transition.
\section{Computational Details}
We used the structural data of BiCoO$_{3}$ measured by the neutron
power diffraction.~\cite{Oka} Our calculations were performed
using the full-potential augmented plane-wave plus local-orbital
code WIEN2k.~\cite{Blaha} The muffin-tin sphere radii were chosen to
be 2.3, 1.9, and 1.4 bohr for Bi, Co, and O atoms, respectively
(1.0 bohr for O when calculating the $E$-$V$ curves shown in Fig. 6).
The cutoff energy of 16 Ryd was set for the plane-wave expansion of
interstitial wave functions, and 1200 {\bf k} points in the first
Brillouin zone for the ambient structure with one formula unit (f.u.) and
300 {\bf k} points for the HP phase with 4 f.u.
\begin{figure}
\includegraphics[angle=0,width=6cm]{Fig1.eps}
\caption{(Color online) $E$ $vs$ $c$/$a$ curve calculated by GGA+$U$ for the Co$^{3+}$-LS
relaxed structures of BiCoO$_3$ with $c$/$a$ = 1-1.3 (in a step of 0.05)
including 1.27 (expt.). The inset shows a unit cell of the experimental
tetragonal structure.}
\label{fgr:1}
\end{figure}
Plain GGA [or local-spin-density approximation (LSDA)]
calculations~\cite{Uratani,Ravindran,Ming,Cai} seem to qualitatively
reproduce the C-type AF and insulating ground state of BiCoO$_3$ in the
ambient phase,
which was ascribed to the strong Hund-exchange stabilized HS state
of the Co$^{3+}$ ions (and thus the AF order and narrow bands) and
to the well split-off $xy$-singlet orbital.
However, the band gap of 0.6 eV and the Co$^{3+}$ spin moment of 2.4 $\mu_B$
given by the GGA/LSDA calculations are much smaller than the experimental
values of 1.7 eV and 3.2 $\mu_B$.\cite{Belik,McLeod}
Moreover, a recent prediction of an insulator-metal transition in
BiCoO$_3$ upon a volume decrease of $\sim$5\% made by GGA/LSDA
calculations~\cite{Ravindran,Ming} has already been disproved by
a very recent HP study,~\cite{Oka} which shows that BiCoO$_3$ is
still semiconducting even under 6 GPa with a large volume decrease of 18\%.
As seen below, the experimental values of both the band gap and the Co$^{3+}$ spin
moment are well reproduced by our GGA+$U$
calculations. Note also that BiCoO$_3$ has an apparent AF order
up to 470 K.\cite{Belik}
All these suggest that BiCoO$_3$ should rather be categorized as a
Mott insulator, with its band gap determined primarily by Hubbard $U$, i.e.,
strong correlation of the Co $3d$ electrons.
To account for the strong electronic correlation,~\cite{Anisimov} we have carried out
GGA+$U$ calculations throughout this paper.
In particular, we used the configuration-state-constrained GGA+$U$
method,~\cite{Korotin,Knizek,Wu1} which allows us to access different
spin and orbital configuration states
of the concern by initializing their corresponding density matrix
and then doing self-consistent electronic relaxation.~\cite{Wu2}
This method is quite useful for study of the spin and orbital
physics present in transition-metal oxides.~\cite{Korotin,Knizek,Wu1,Wu2} All the
results shown below are obtained with $U$=6 eV and Hund exchange
$J$=0.9 eV. We note that our test calculations using other
reasonable $U$ values,~\cite{Cai,Wu1,Wu2} $U$=5 and 7 eV, gave
qualitatively the same results.
\section{Results and Discussion}
We first seek the origin of the giant tetragonal distortion
by starting with our calculations assuming an ideal cubic structure
(space group $Pm\bar{3}m$) having the same volume as the experimental ambient
tetragonal structure. Then we changed the $c$/$a$ ratio of the tetragonal structure
(keeping the volume unchanged) in our calculations. All those calculations
were carried out by setting a hypothetical LS state of the Co$^{3+}$ ions, and
by doing a full electronic and atomic relaxation for the cases of
$c$/$a$ = 1.05-1.3 in a step of 0.05, and for the experimental 1.27 as well.
All the solutions are insulating.
As the LS Co$^{3+}$ has a closed sub-shell $t_{2g}^6$
(thus no orbital degeneracy) and is an isotropic ion, we can use this
set of LS-constrained GGA+$U$ calculations as a computer experiment to probe mainly
the Bi-O covalent effect.
\begin{table}
\caption{The total energies of BiCoO$_3$
(in unit of eV/f.u.) relative to the hypothetical cubic-structure LS state
calculated by GGA+$U$:
the experimental tetragonal phase assuming an LS state with atomic relaxations
(the third row),
assuming an HS state in the LS relaxed structure (the fourth row), and assuming the HS state
with further atomic relaxations (the fifth (last) row). The optimized atomic $z$-coordinates are also shown.
Note that the structural data
of the atomically relaxed HS state are in good agreement with the
experiment~\cite{Oka} (see the inset of Fig. 1).
See more discussion in the main text.}
\label{tbl:1}
\begin{tabular}{lcc} \hline
States & $\Delta$$E$ & $z_{\rm Co}$, $z_{\rm O1}$, $z_{\rm O2}$ \\ \hline
Cubic LS & 0 & \\
Tetragonal LS-relaxed & --0.90 & 0.5897, 0.1889, 0.6853 \\
Tetragonal HS at LS-relaxed & --1.47 & \\
Tetragonal HS-relaxed & --1.97 & 0.5631, 0.1912, 0.7237 \\
\hline
\end{tabular}
\end{table}
We show in Fig. 1 the calculated total energies as function
of the $c$/$a$ ratios, and one can immediately find that the $c$/$a$ ratio in the
LS equilibrium state is close to 1.2 (about 1.18), already indicating a large
tetragonal distortion. The corresponding energy gain, relative to the hypothetical
cubic structure ($c$/$a$ = 1), is 0.96 eV/f.u. For the LS relaxed structure with the
experimental $c$/$a$ = 1.27, the energy gain is 0.90 eV/f.u.
We list in Table I (see the third row) the optimized atomic $z$-coordinates of the
LS state with the experimental $c$/$a$ ratio,
and we find that the experimental CoO$_5$ coordination is well achieved
[as seen from the optimized Co-O bondlengths, Co-O1: 1.893 \AA$\times$1,
2.832 \AA$\times$1 (much larger); Co-O2: 1.919 \AA$\times$4]
even in the presence of the Co$^{3+}$ LS state without orbital degeneracy.
This is also the case for $c$/$a$ = 1.2 with another optimized Co-O bondlengths,
1.881 \AA$\times$1, 2.677 \AA$\times$1, and 1.943 \AA$\times$4.
We plot in Fig. 2(a) a charge density contour of the hypothetical cubic structure
in the (100) plane, and in Fig. 2(b) that of the LS relaxed state with the
experimental $c$/$a$ ratio. Fig. 2(a) shows nearly spherical charge densities
around both the Bi and O ions, indicative of Bi$^{3+}$-O$^{2-}$ ionic bonds
in the hypothetical cubic structure. In contrast, Fig. 2(b) clearly shows directional
Bi-O covalent bonds, which are also indicated the orbitally resolved density of
states (see Fig. 3 and the discussion below).
All these suggest that the Bi-O covalency drives the giant tetragonal
FE distortion
and causes remarkable off-center atomic displacements, no matter that
the Co$^{3+}$ ions are in
this hypothetical LS state or in the real HS state (see below).
This is similar to the stereochemical mechanism of the
Bi$^{3+}$ 6s$^2$ lone-pairs proposed for the highly distorted perovskite
manganite BiMnO$_3$.~\cite{Seshadri}
\begin{figure}[h]
\includegraphics[angle=270,width=7cm]{Fig2.eps}
\caption{(Color online) Charge density contour (0.01-0.1 $e$/\AA$^3$ in a step of 0.01 $e$/\AA$^3$)
of BiCoO$_3$ in the
(100) plane of (a) the hypothetical ideal cubic structure and of (b) the
LS-relaxed structure with the experimental $c$/$a$ ratio.
The directional Bi-O covalency is apparent in (b).}
\label{fgr:2}
\end{figure}
\begin{figure}[h]
\includegraphics[angle=0,width=5.5cm]{Fig3.eps}
\caption{Partial density of states of BiCoO$_3$ in the C-type AF ground state of the
ambient phase.}
\label{fgr:3}
\end{figure}
Furthermore, we study the effect of the $xy$-type OO of the HS Co$^{3+}$ ions
on the giant tetragonal distortion.
Owing to the effective
CoO$_5$ pyramidal coordination already present in the LS relaxed tetragonal structures
as described above,
the crystal field (particularly the $t_{2g}$-$e_g$ splitting) is weak and thus
less important than the Hund
exchange.~\cite{Hu} As a result, an HS state with an OO of the well
split-off $xy$ orbital
is the ground state and more stable than the hypothetical LS state by 0.57 eV/f.u. for the
experimental $c$/$a$ = 1.27 (see the third and fourth rows in Table I) and by 0.59 eV/f.u.
for $c$/$a$ = 1.2.
In this sense, the $xy$ OO is a consequence of the Bi-O covalency
driven tetragonal distortion and the associated CoO$_5$ pyramidal coordination.
Moreover, our calculations doing structural optimization show that an adjustment of
the lattice to this $xy$ OO of the HS Co$^{3+}$ ions
changes the $c$/$a$ ratio to 1.28 (as compared to 1.18 for the LS equilibrium state,
see Fig. 1). This agrees very well with the experimental $c$/$a$ = 1.27.
Correspondingly, the experimental atomic parameters are also well reproduced, see the last row
in Table I. Furthermore, we show in Fig. 3 the density of states of BiCoO$_3$ in the
C-type AF ground state having the $xy$ OO of the HS Co$^{3+}$ ions. The calculated band gap
of 1.98 eV and the Co$^{3+}$ spin moment of 3.01 $\mu_B$ are also in good agreement with
the experimental values of 1.7 eV and 3.24 $\mu_B$.\cite{Belik,McLeod}
Hybridizations between the Bi $6s6p$
and O $2p$ orbitals are also apparent and evidence again the Bi-O covalency,
although the magnitude of their respective
density of states is much underestimated within the muffin-tin spheres as those orbitals
are spatially quite spread. In a word, all above results allow us to conclude that
the giant
tetragonal FE distortion of BiCoO$_3$ originates from the Bi-O covalency
(rather than from lifting of the orbital degeneracy of the HS Co$^{3+}$
ions~\cite{Uratani,Oka,Okuno}) and is further stabilized by the subsequent $xy$-type OO.
We now identify the nature of the pressure induced spin-state
transition in BiCoO$_3$, using another set of constrained
GGA+$U$ calculations for the orthorhombic structure measured at 5.8 GPa, by assuming
the LS, IS, and the mixed HS+LS states, respectively, and by doing a full electronic
and atomic relaxation for each case.
\begin{table}
\caption{The relative total energies $\Delta$$E$ (meV/f.u.), Co$^{3+}$ spin
moments $M$ ($\mu$$_{B}$), and band gap \emph{E$_{g}$} (eV) of
BiCoO$_{3}$ in the 5.8 GPa phase calculated by GGA+$U$ for
the LS-relaxed, IS-relaxed, and (HS+LS)-relaxed structures.}
\label{tbl:2}
\begin{tabular}{lccc}
\hline
States&$\Delta$$E$&$M_{\rm Co}$&\emph{E$_{g}$}\\
\hline
(HS+LS)-relaxed&0&3.07, 0.17&0.84\\
LS-relaxed&91&0&0.95\\
IS-relaxed&158&2.01&half-metal\\
\hline
\end{tabular}
\end{table}
\begin{figure}[h]
\includegraphics[angle=0,width=6cm]{Fig4.eps}
\caption{(Color online) The total and Co $3d$ density of states of BiCoO$_3$ in the 5.8 GPa phase
calculated by GGA+$U$ for the relaxed IS (a), LS (b), and mixed HS+LS (c) structures,
respectively.}
\label{fgr:4}
\end{figure}
As seen in Table II, the mixed HS+LS state has the lowest total energy, and the LS (IS)
state lies above it by 91 (158) meV/f.u. Those results suggest that
either the pure LS or IS state present in the 5.8 GPa phase~\cite{Oka} is not the case.
Reversely, if the pure LS state were present in the 5.8 GPa phase,
it would give rise to a change of the spin state from the pure HS state in
the ambient phase, $\Delta$$S$=2, being in disagreement with the
observed $\Delta$$S$=1.\cite{Oka} Moreover, an absence of the IS state is also not
surprising, as (1) BiCoO$_3$ in
the 5.8 GPa phase is free of a Jahn-Teller distortion (that is expected for the localized IS
Co$^{3+}$);
(2) a half-metallic band structure of the IS state (see Fig. 4(a))
disagrees with the measured semiconducting
behavior;~\cite{Oka} and (3) up to now a definite example of the insulating IS state appears
still lacking, and even in the layered perovskites LaSrCoO$_4$ and
La$_{1.5}$Sr$_{0.5}$CoO$_4$ both
having a strong tetragonal elongation of the Co$^{3+}$O$_6$ octahedra, the IS state
turns out not to be the ground state either,~\cite{Wu2} despite an IS state
might be intuitively expected.
As such, the mixed HS+LS state is most probably present in the 5.8 GPa phase.
An ideal 1:1 configuration of the mixed HS+LS state, with an average $S$=1, well
accounts for the observed change of the spin state.~\cite{Oka}
Moreover, a G-type order of the HS and LS Co$^{3+}$ ions (each HS Co
ion is surrounded by six LS Co ions, and vice versa), due to a
bigger/smaller size of the HS/LS Co$^{3+}$ ions, could help to gain
an elastic energy, and the resultant Co-O bondlengths are calculated to be
1.980 (1.930), 1.976 (1.890), and 1.948 (1.906) \AA~ for the HS (LS)
Co$^{3+}$ ions along the local $xyz$ axes.
Note, however, that a long-range G-type
order of the HS and LS Co$^{3+}$ ions is hard to establish, as only
single transition-metal species in a single charge state and in the
identical octahedral coordinations, i.e., solely Co$^{3+}$ ions are involved.
Thus, the average Co-O bondlengths of the mixed and disordered HS+LS state
(possibly with a short-range order due to a partial release of the lattice elasticity)
are also in good agreement with
the experiment.~\cite{Oka} Furthermore, the calculated insulating gap of
BiCoO$_3$ is reduced from 1.98 eV in the C-type AF state of the
ambient phase to 0.84 eV in the mixed HS+LS state (see Fig. 4(c)), which
qualitatively accounts for the decreasing resistivity of BiCoO$_3$
under pressure.~\cite{Oka}
\begin{figure}
\includegraphics[angle=0,width=6cm]{Fig5.eps}
\caption{Electron hopping (a) from an HS Co$^{2+}$ ion
to a neighboring HS Co$^{3+}$ and (c) from an LS Co$^{3+}$ to
an LS Co$^{4+}$, but a suppressed electron hopping (b)
from an HS Co$^{3+}$ to an LS Co$^{4+}$ due to a spin blockade.}
\label{fgr:5}
\end{figure}
Note that when considering thermal
excitation of electrons into the initially empty conduction band and
holes left in the valence band for a nominally stoichiometric
material, these electron excitations would in a localized picture
correspond to HS Co$^{2+}$ states and holes to LS Co$^{4+}$.
They behave like HS Co$^{2+}$ and LS Co$^{4+}$ ``impurities" in the matrix
of the Co$^{3+}$ ions.
As a result, in the ambient phase of BiCoO$_3$ having the HS Co$^{3+}$ matrix,
only the HS Co$^{2+}$ ``impurities" can transfer their
minority-spin $t_{2g}$ electrons to the neighboring HS Co$^{3+}$,
without changing the configuration states (the initial and
final states are the same), see Fig. 5(a). However, a hole hopping from
the LS Co$^{4+}$ to the HS Co$^{3+}$ is significantly
suppressed (see Fig. 5(b)), due to a cost of the Hund exchange energy associated
with a large change of the spin states which is referred as to a
spin-blockade machanism.~\cite{Maignan} In contrast, in the HP phase having
the mixed and disordered HS+LS Co$^{3+}$ matrix,
a charge hopping can take place both
between the HS Co$^{2+}$ and HS Co$^{3+}$ (Fig. 5(a)) and between the LS
Co$^{4+}$ and LS Co$^{3+}$ (Fig. 5(c)). This could also account for the
decreasing resistivity of BiCoO$_3$ under pressure.
Our above results show that
even in the 5.8 GPa phase of BiCoO$_3$ with a large volume decrease of 18\%,
the HS-to-LS
transition is not yet complete, and the system is most probably in the mixed HS+LS
insulating state, but not in the pure-LS metallic state which was predicted by
the previous GGA/LSDA calculations for BiCoO$_3$ upon a volume decrease of
$\ge$5\%.~\cite{Ravindran,Ming}
As we discussed above, BiCoO$_3$ is rather a Mott insulator,
and thus the previous GGA/LSDA results are somewhat questionable:
particularly the predicted
metallic solution was already disproved by a very recent experiment.~\cite{Oka}
Recent fixed-spin-moment calculations even predicted that BiCoO$_3$ could have a giant
magnetoelectric coupling with a readily switchable HS-LS transition
associated with an electric field driven FE-PE transition.~\cite{Ravindran}
We note that
the prediction may simply be an artifact of the fixed-spin-moment calculations,
as (1) LSDA or GGA (it was mentioned as a density-functional method in
Ref~\cite{Ravindran})
is not well suitable to describe this
Mott insulator; (2) most probably those calculations were carried out in a wrong
ferromagnetic metallic state;
and (3) their metallic solutions blurred the distinction between
the different spin and orbital multiplets of the concern and thus suppressed
significantly their level
splittings and particularly the HS/LS splitting: the fixed-spin-moment
calculations showed that
for the ambient structure, the energy preference of the HS state over the LS state is
less than 0.15 eV/f.u. (see Fig. 2 in Ref~\cite{Ravindran}),
whereas the corresponding value we calculated is more than 0.5 eV/f.u. Therefore, the previous
prediction of the HS-to-LS transition in BiCoO$_3$ with a small volume decrease of $\sim$5\%
was overly optimistic.~\cite{Ravindran} Instead, we find now that there is no readily switchable HS-LS
transition is BiCoO$_3$.
\begin{figure}
\includegraphics[angle=0,width=6cm]{Fig6.eps}
\caption{$E$-$V$ curves calculated by GGA+$U$ for the C-type AF state of BiCoO$_3$
in the ambient phase, and for the mixed HS+LS state and the LS state under high pressures.
The lattice volume of the ambient phase is slightly overestimated within 3\%, and
the critical volume for a complete transition into the LS state is estimated to be
about 53 \AA$^{3}$/f.u.}
\label{fgr:6}
\end{figure}
As Bi$^{3+}$ has a very similar ionic size as La$^{3+}$,
it is reasonable to assume that when the local Co-O bondlengths of BiCoO$_3$
in the HP PE phase become identical to those of LaCoO$_3$ in the LS state at low
temperature (1.925 \AA~ at 5 K),~\cite{Radaelli}
a complete transition into the LS state would be achieved in BiCoO$_3$.
This, together with the structural data of the 5.8 GPa phase,\cite{Oka}
allows us to estimate the critical volume to be about 52.8 \AA$^{3}$/f.u.
(corresponding to a volume decrease of about 20\%).
Then, by extrapolating the eye-guided line of the $V$-$P$ data points
in the range of 2-6 GPa (see Fig. 1(b) in Ref~\cite{Oka}), we may estimate the
critical pressure to be about 8 GPa.
It is important to note that the estimated critical volume is indeed well reproduced
by our detailed calculations of the $E$-$V$ curves
(see Fig. 6, $V_{\rm c}\approx$ 53 \AA$^{3}$/f.u.),
which also nicely reproduce the equilibrium volume of the ambient phase within 3\%
and clearly indicate the HS-to-LS transition via the mixed HS+LS state.
This prediction of a complete transition into the LS state
awaits a further HP study.
\section{Conclusion}
To conclude, using configuration-state-constrained GGA+$U$ calculations,
we demonstrate that the giant tetragonal ferroelectric distortion of BiCoO$_3$ is driven
by the strong Bi-O covalency (rather than by lifting of the
orbital degeneracy of the HS Co$^{3+}$ ions) and is further stabilized by
a subsequent $xy$-type OO.
Moreover, our results show that the pressure induced HS-to-LS transition
is via a mixed HS+LS state,
and that the transition would be complete upon a large volume decrease of
about 20\% (under about 8 GPa). The mixed HS+LS state well accounts for the
available experimental results.\cite{Oka}
\section{Acknowledgments}
H. Wu is supported by the DFG via SFB 608 and by Fudan Univ.
The research at Hefei is funded by the NSF of China (Grant No. 11004195),
the Special Funds for Major State Basic Research Project of China (973)
under Grant No. 2007CB925004, Knowledge Innovation Program of CAS
(Grant No. KJCX2-YW-W07), Director Grants of CASHIPS, and CUHK
(Grant No. 3110023).
|
\section*{ Introduction}
\medskip
In the given article the notion of infinite order decomposition of
a C*-algebra with respect to an infinite orthogonal set of
projections is investigated. It is known that for any projection
$p$ of a C$^*$-algebra $A$ the next equality is valid $A=pAp\oplus
pA(1-p)\oplus (1-p)Ap \oplus (1-p)A(1-p)$, where $\oplus$ is a
direct sum of spaces. In the given article we investigated an
infinite analog of this decomposition, an infinite order
decomposition. The notion of infinite order decomposition was
introduced in \cite{1}. The next theorems belong to \cite{1}:
let $A$ be a C$^*$-algebra of on a Hilbert space $H$, $\{p_\xi\}$
be an infinite orthogonal set of projections in $A$ with the least
upper bound $1$ in the algebra $B(H)$. Then
{\it 1) if the order unit space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is monotone complete in $B(H)$ (i.e. ultraweakly closed),
then $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a C$^*$-algebra.
2) if $A$ is monotone complete in $B(H)$ (i.e. a von Neumann
algebra), then $A=\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$.}
In the given article we proved that for the infinite order
decomposition $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ of a
C$^*$-algebra $A$ with respect to an infinite orthogonal set
$\{p_i\}$ of projections of $A$, if $p_\xi Ap_\xi$ is a von
Neumann algebra for any $\xi$ then $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is a von Neumann algebra. For this propose it was
constructed a multiplication and an involution corresponding to
infinite order decompositions. It turns out, the order and the
norm defined in the infinite order decomposition of a
C$^*$-algebra on a Hilbert space $H$ coincide with the usual order
and the norm in the algebra $B(H)$. Also, it is proved that, if a
C$^*$-algebra $A$ with an infinite orthogonal set $\{p_\xi\}$ of
projections in $A$ such that $\sup_\xi p_\xi=1$ is not a von
Nemann algebra, projections of the set $\{p_\xi\}$ are pearwise
equivalent then $A\neq \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$.
Moreover if the order unit space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is not weakly closed then $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is not a C$^*$-algebra.
The author wants to thank A.A.Rakhimov for many stimulating
conversations on the subject.
\newpage
\bigskip
\section{Infinite order decompositions}
\medskip
Let $A$ be a C$^*$-algebra on a Hilbert space $H$, $\{p_\xi\}$ be
an infinite orthogonal set of projections of the algebra $A$ with
the least upper bound $1$ in the algebra $A$. By
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ we denote the set
$$
\{\{a_{\xi,\eta}\}: a_{\xi,\eta}\in
p_\xi Ap_\eta\,\, for\, all \,\,\xi, \eta, \,and\,there\, exists\,
such\, number
$$
$$
K\in R \,\,that \Vert\sum_{k,l=1}^na_{kl}\Vert\le K \,\,for\,
all\,\, n\in N \,\,and\,\,\,\{a_{kl}\}_{kl=1}^n\subseteq
\{a_{\xi,\eta}\}\},
$$
and say $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is {\it an infinite
order decomposition} of the algebra $A$.
Let $A$ be a C$^*$-algebra on a Hilbert space $H$, $\{p_\xi\}$ be
an infinite orthogonal set of projections of the algebra $A$ with
the least upper bound $1$ in the algebra $B(H)$. We define a
relation of an order $\leq$ in the vector space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ as follows: for elements
$\{a_{\xi\eta}\}$, $\{b_{\xi\eta}\}\in \sum_{\xi,\eta}^\oplus
p_\xi Ap_\eta$, if for all $n\in N$, $\{p_k\}_{k=1}^n\subset
\{p_\xi\}$ the inequality $\sum_{k,l=1}^na_{kl}\leq
\sum_{k,l=1}^nb_{kl}$ holds, then we will write
$\{a_{\xi\eta}\}\leq \{b_{\xi\eta}\}$. Also, the map
$\{a_{\xi,\eta}\}\to\Vert \{a_{\xi,\eta}\}\Vert$,
$\{a_{\xi,\eta}\}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, where
$\Vert \{a_{\xi,\eta}\}\Vert=\sup \{\Vert \sum_{kl=1}^n
a_{kl}\Vert :n\in N, \{a_{kl}\}_{kl=1}^n \subseteq
\{a_{\xi,\eta}\}\}$, is a norm on vector space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$.
{\it Example.} Let $n$ be an arbitrary infinite cardinal number,
$\Xi$ be a set of indexes of the cardinality $n$. Let $\{e_{ij}\}$
be a set of matrix units such that $e_{ij}$ is a $n\times
n$-dimensional matrix, i.e.
$e_{ij}=(a_{\alpha\beta})_{\alpha\beta\in\Xi}$, the $(i,j)$-th
component of which is $1$, i.e. $a_{ij}=1$, and the rest
components are zeros. Let $\{m_\xi\}_{\xi\in \Xi}$ be a set of
$n\times n$-dimensional matrixes. By $\sum_{\xi\in \Xi} m_\xi$ we
denote the matrix whose components are sums of the corresponding
components of matrixes of the set $\{m_\xi \}_{\xi\in \Xi}$. Let
$$
M_n({\bf C})=\{\{\lambda_{ij}e_{ij}\}: \,for\,\, all\,\,
indexes\,\, i,\,j \,\lambda_{ij}\in {\bf C},
$$
$$
and\,\, there\,\, exists\,\, such\,\, number\,\, K\in {\bf
R},\,\,that \,\, for \,\, all\,\, n\in N
$$
$$
and\,\, \{e_{kl}\}_{kl=1}^n\subseteq \{e_{ij}\} \Vert\sum_{kl=1}^n
\lambda_{kl}e_{kl}\Vert \le K\},
$$
where $\Vert \,\, \Vert$ is a norm of a matrix. It is easy to see
that $M_n({\bf C})$ is a vector space. The set $M_n({\bf C})$,
defined above, coincides with the next set:
$$
\mathcal{M}_n({\bf C})=\{\{\lambda_{ij}e_{ij}\}:\,for\,\, all\,\,
indexes\, ij\,\,\lambda_{ij}\in {\bf C},
$$
$$
and\,\,\, there\,\,\, exists\,\,\, such\,\,\, number\,\,\, K\in R
\,\,that \,\, for\,\, all\,
$$
$$
\{x_i\}\in l_2(\Xi)\,\, the\,\,\, next\,\,\, inequality\,\,\,
holds \, \sum_{j\in \Xi} \vert\sum_{i\in \Xi}\lambda_{ij}x_i
\vert^2\le K^2\sum_{i\in \Xi} \vert x_i\vert^2\},
$$
where $l_2(\Xi)$ is a Hilbert space on ${\bf C}$ with elements
$\{x_i\}_{i\in \Xi}$, where $x_i\in \bf C$ for all $i\in \Xi$.
The associative multiplication of elements in $M_n({\bf C})$ can
be defined as follows: if $x=\sum_{ij\in \Xi}\lambda_{ij}e_{ij}$,
$ y=\sum_{ij\in \Xi}\mu_{ij}e_{ij}$ are elements of $M_n({\bf C})$
then $xy=\sum_{ij\in \Xi} \sum_{\xi\in \Xi} \lambda_{i\xi}\mu_{\xi
j}e_{ij}$. On this operation $M_n({\bf C})$ is an associative
algebra and $M_n({\bf C})=B(l_2(\Xi))$, where $B(l_2(\Xi))$ is the
associative algebra of all bounded linear operators on the Hilbert
space $l_2(\Xi)$. Then $M_n({\bf C})$ is a von Neumann algebra of
infinite $n\times n$-dimensional matrixes on ${\bf C}$, who is
defined by its own infinite order decomposition.
Analogously, if we take the algebra $B(H)$ of all bounded linear
operators on an arbitrary Hilbert space $H$ and if $\{q_i\}$ is an
arbitrary maximal orthogonal set of minimal projections of the
algebra $B(H)$, then $B(H)=\sum_{ij}^\oplus q_i B(H)q_j$ (see
\cite{1}).
\bigskip
Let $A$ be a C$^*$-algebra, $\{p_i\}$ be an infinite orthogonal
set of projections with the least upper bound $1$ in the algebra
$A$ and let $\mathcal{A}=\{ \{p_iap_j\}: a\in A\}$. Then $A\equiv
\mathcal{A}$ (see \cite{2}).
{\bf Lemma 1.} {\it Let $A$ be a C$^*$-algebra, $\{p_\xi\}$ be an
infinite orthogonal set of projections of the algebra $A$ with the
least upper bound $1$ in the algebra $A$. Then,
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a vector space with the
next componentwise algebraic operations
$$
\lambda\cdot \{a_{\xi\eta}\}=\{ \lambda a_{\xi\eta}\}, \lambda\in
{\bf C}
$$
$$
\{a_{\xi\eta}\}+\{b_{\xi\eta}\}=\{a_{\xi\eta}+b_{\xi\eta}\},
a_{\xi\eta}, b_{\xi\eta}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta .
$$
And the space $\mathcal{A}$ is a vector subspace of the vector
space $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. }
\medskip
{\bf Lemma 2.} {\it Let $A$ be a C$^*$-algebra, $\{p_\xi\}$ be an
infinite orthogonal set of projections of the algebra $A$ with the
least upper bound $1$ in the algebra $A$. Then, the map
$\{a_{\xi,\eta}\}\to\Vert \{a_{\xi,\eta}\}\Vert$,
$\{a_{\xi,\eta}\}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, where
$\Vert \{a_{\xi,\eta}\}\Vert=\sup \{\Vert \sum_{kl=1}^n
a_{kl}\Vert :n\in N, \{a_{kl}\}_{kl=1}^n \subseteq
\{a_{\xi,\eta}\}\}$, is a norm, and $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is a Banach space with this norm.}
{\bf Proof.} It is clear, that for any element
$\{a_{\xi,\eta}\}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, if
$\Vert \{a_{\xi,\eta}\}\Vert=0$, then $a_{\xi,\eta}=0$ for all
$\xi$, $\eta$, i.e. $\{a_{\xi,\eta}\}=0$. The rest conditions in
the definition of the norm also can be easily checked.
Let $(a_n)$ be an arbitrary Cauchy sequence in the space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, i.e. for any positive
number $\varepsilon >0$ there exists $n\in {\bf N}$ such, that for
all $n_1\geq n$, $n_2\geq n$ the inequality $\Vert
a_{n_1}-a_{n_2}\Vert<\varepsilon$ holds. Then the set $\{\Vert
a_n\Vert\}$ is bounded by some number $K\in {\bf R}_+$ and for any
finite set $\{p_k\}_{k=1}^n\subset \{p_i\}$ the sequence $(pa_np)$
is a Cauchy sequence, where $p=\sum_{k=1}^n p_k$. Then, since $A$
is a Banach space, then $\lim_{n\to \infty} pa_np\in A$.
Let $a_{\xi,\eta}=\lim_{n\to \infty} p_\xi a_np_\eta$ for all
$\xi$ and $\eta$. Then $\Vert \sum_{kl=1}^n a_{kl}\Vert\leq K$ for
all $n\in {\bf N}$ and $\{a_{kl}\}_{kl=1}^n \subseteq
\{a_{\xi,\eta}\}$. Hence $\{a_{\xi,\eta}\}\in
\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. $\triangleright$
\medskip
The definition of the order in $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is equivalent to the next condition: for the elements
$\{a_{\xi\eta}\}$, $\{b_{\xi\eta}\}\in\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$, if for all $n\in N$ and $\{p_k\}_{k=1}^n \subseteq
\{p_i\}$ the equality $\{a_{kl}\}_{k,l=1}^n\leq
\{b_{kl}\}_{k,l=1}^n$ holds in the algebra $\mathcal{A}$, then
$\{a_{\xi\eta}\}\leq \{b_{\xi\eta}\}$.
{\bf Proposition 3.} {\it Let $A$ be a C$^*$-algebra on a Hilbert
space $H$, $\{p_\xi\}$ be an infinite orthogonal set of
projections in $A$ with the least upper bound $1$ in the algebra
$B(H)$. Then the relation $\leq$, introduced above, is a relation
of a partial order, and the space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is an order unit space with this order. In this case
$\mathcal{A}=\{\{p_\xi ap_\eta\}: a\in A\}$ is an order unit
subspace of the order unit space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$. }
{\bf Proof.} Let $\mathcal{M}=\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$. The space $\mathcal{M}$ is a partially ordered vector
space, i.e. $\mathcal{M}_+\cap \mathcal{M}_-=\{0\}$, where
$\mathcal{M}_+=\{ \{a_{\xi\eta}\}\in \mathcal{M}:
\{a_{\xi\eta}\}\geq 0\}$, $\mathcal{M}_-=\{ \{a_{\xi\eta}\}\in
\mathcal{M}: \{a_{\xi\eta}\}\leq 0\}$.
By the definition of the order the partially ordered vector space
$\mathcal{M}$ is Archimedean. Let $\{a_{\xi\eta}\}\in
\mathcal{M}$. Since for any finite set $\{p_k\}_{k=1}^n\subset
\{p_\xi\}$ the inequality $-\Vert \{a_{\xi,\eta}\}\Vert p\leq
p\{a_{\xi,\eta}\}p\leq \Vert \{a_{\xi,\eta}\}\Vert p$ holds, where
$p=\sum_{k=1}^n p_k$, then by the definition of the order $-\Vert
\{a_{\xi,\eta}\}\Vert 1\leq \{a_{\xi,\eta}\}\leq \Vert
\{a_{\xi,\eta}\}\Vert 1$, and the unit of $A$ is an order unit of
the partially ordered vector space $\mathcal{M}$. Thus,
$\mathcal{M}$ is an order unit space.
By lemma 1 $\mathcal{A}$ is an order unit subspace of the order
unit space $\mathcal{M}$. $\triangleright$
\medskip
{\bf Proposition 4.} {\it Let $A$ be a C$^*$-algebra on a Hilbert
space $H$, $\{p_i\}$ be an infinite orthogonal set of projections
in $A$ with the least upper bound $1$ in the algebra $B(H)$. Then
the order unit space $\mathcal{A}=\{\{p_\xi ap_\eta\}: a\in A\}$
is a C$^*$-algebra, where the operation of multiplication of
$\mathcal{A}$ defines as follows
$$
\cdot :<\{p_\xi ap_\eta\},\{p_\xi bp_\eta\}>\to \{p_\xi
abp_\eta\}, \{p_\xi ap_\eta\},\{p_\xi bp_\eta\}\in \mathcal{A}.
$$
}
{\bf Proof.} By lemma 4 in \cite{2} the map
$$
\mathcal{I}: a\in A\to \{p_\xi ap_\eta\}\in \mathcal{A}
$$
is a one-to-one map. In this case
$$
\mathcal{I}(a)\mathcal{I}(b)=\mathcal{I}(ab)
$$
by the definition in proposition 4 of the multiplication, and
$\mathcal{I}(a)=\{p_\xi ap_\eta\}$, $\mathcal{I}(b)=\{p_\xi
bp_\eta\}$, $\mathcal{I}(ab)=\{p_\xi ab p_\eta\}$. Hence, the
operation, introduced in the formulation of proposition 4 is an
associative multiplication and the map $\mathcal{I}$ is an
isomorphism of the algebras $A$ and $\mathcal{A}$.
By proposition 3 the isomorphism $\mathcal{I}$ is isometrical.
Therefore $\mathcal{A}$ is a C$^*$-algebra with this operation.
$\triangleright$
\medskip
{\it Example 1.} We take the algebra $B(H)$ of all bounded linear
operators on a Hilbert space $H$. Let $\{q_i\}$ be a maximal
orthogonal set of minimal projections of the algebra $B(H)$. Then
$\sup_i q_i=1$ and by lemma 4 in \cite{2} and proposition 4 the
space $\mathcal{B(H)}=\{\{q_iaq_j\}: a\in B(H)\}$ can be
identified with $B(H)$ as C$^*$-algebras in the sense of the map
$$
\mathcal{I}: a\in B(H)\to \{q_iaq_j\}\in \mathcal{B(H)}.
$$
In this case the operation associative multiplication in
$\mathcal{B(H)}$ is defined as follows
$$
\cdot :<\{q_i aq_j\},\{q_i bq_j\}>\to \{q_i abq_j\}, \{q_i
aq_j\},\{q_i bq_j\}\in \mathcal{B(H)}.
$$
Let $a$, $b\in B(H)$, $q_iaq_j=\lambda_{ij}q_{ij}$,
$q_ibq_j=\mu_{ij}q_{ij}$, where $\lambda_{ij}$, $\mu_{ij}\in {\bf
C}$, $q_i=q_{ij}q_{ij}^*$, $q_j=q_{ij}^*q_{ij}$, for all indexes
$i$ and $j$. Then this multiplication coincides with the next
bilinear operation
$$
\cdot :<\{q_i aq_j\},\{q_i bq_j\}>\to \{\sum_\xi \lambda_{i\xi}
\mu_{\xi j}q_{ij}\}, \{q_i aq_j\},\{q_i bq_j\}\in \mathcal{B(H)}.
$$
\medskip
{\it Remark 1.} Let $A$ be a C$^*$-algebra on a Hilbert space
$H$, $\{p_i\}$ be an infinite orthogonal set of projections in $A$
with the least upper bound $1$ in the algebra $B(H)$. Then by
proposition 4 $\mathcal{A}=\{\{p_\xi ap_\eta\}: a\in A\}$ is a
C$^*$-algebra. In this case the involution on the algebra
$\mathcal{A}$ coincides with the next map:
$$
\{p_\xi ap_\eta\}^*=\{p_\xi a^*p_\eta\},\,\,a\in A.
$$
Indeed, the identification $\mathcal{A}\equiv A$ gives us
$a=\{p_\xi ap_\eta\}$ and $a^*=\{p_\xi a^*p_\eta\}$ for all $a\in
A$. Then $\{p_\xi ap_\eta\}^*=a^*=\{p_\xi a^*p_\eta\}$ for any
$a\in A$. Let $\mathcal{A}_{sa}=\{\{p_\xi ap_\eta\}: a\in
A_{sa}\}$. Then $\mathcal{A}=\mathcal{A}_{sa}+i\mathcal{A}_{sa}$.
Indeed, $\{p_\xi ap_\eta\}^*=a^*=a=\{p_\xi ap_\eta\}$ for any
$a\in A_{sa}$.
Let $\mathcal{N}=\{\{p_\xi ap_\eta\}: a\in B(H)\}$. By lemma 4 in
\cite{2} and by proposition 4 $\mathcal{N}\equiv B(H)$. Therefore
we will assume that $\mathcal{N}=B(H)$. Let
$\mathcal{N}_{sa}=\{\{p_\xi ap_\eta\}: a\in B(H), \{p_\xi
ap_\eta\}^*=\{p_\xi ap_\eta\}\}$. Then
$\mathcal{N}=\mathcal{N}_{sa}+i\mathcal{N}_{sa}$. Note that
$\{p_\xi ap_\eta\}^*=\{p_\xi ap_\eta\}$ if and only if $(p_\xi
ap_\eta)^*=p_\eta ap_\xi$ for all $\xi$, $\eta$.
\medskip
{\bf Lemma 5.} {\it Let $B(H)$ be the algebra of all bounded
linear operators on a Hilbert space $H$. Let $\{p_\xi\}$ be an
infinite orthogonal set of projections of the algebra $B(H)$ with
the least upper bound 1. Then the associative multiplication of
the algebra $\mathcal{N}$ (hence of the algebra $B(H)$) coincides
with the next operation
$$
\{p_\xi ap_\eta\}\star \{p_\xi bp_\eta\}=\{\sum_i p_\xi ap_i p_i
bp_\eta\}, \{p_\xi ap_\eta\},\{p_\xi bp_\eta\}\in \mathcal{N}
$$
where the sum $\sum$ in the right part of the equality is an
ultraweak limit of the net of finite sums of elements in the set
$\{p_\xi ap_i p_i bp_\eta\}_{\xi\eta}$. }
{\bf Proof.} Let $\{p_k\}_{k=1}^n$ be a finite subset of the set
$\{p_\xi\}$. Note that $\sup_i p_i=1$ in the algebra $B(H)$, i.e.
the net of all finite sums of the kind $\sum_{k=1}^n p_k$ of
orthogonal projections of the set $\{p_\xi\}$ ultraweakly
converges to the identity operator in $B(H)$. By the ultraweakly
continuity of the operator of multiplication $T(b)=ab, b\in B(H)$,
where $a\in B(H)$, the net of finite sums of elements in the set
$\{p_\xi ap_i p_i bp_\eta\}_{\xi\eta}$ ultraweakly converges in
$B(H)$ and $\sum_i p_\xi ap_i p_i bp_\eta=p_\xi abp_\eta$ for all
$\xi$, $\eta$. Hence the operation of multiplication $\star$ of
the algebra $\mathcal{N}$ coincides with the operation, introduced
in proposition 4. And the operation of the associative
multiplication, introduced in proposition 4 coincides with the
multiplication in the algebra $B(H)$ in the sense
$\mathcal{N}\equiv B(H)$. $\triangleright$
\medskip
{\bf Proposition 6.} {\it Let $A$ be a C$^*$-algebra on a Hilbert
space $H$, $\{p_\xi\}$ be an infinite orthogonal set of
projections in $A$ with the least upper bound $1$ in the algebra
$B(H)$. Then the operation of associative multiplication of the
algebra $\mathcal{A}$ coincides with the inducing on $\mathcal{N}$
of the operation, defined in lemma 5.}
{\bf Proof.} Let $\{p_\xi ap_\eta\}$, $\{p_\xi bp_\eta\}$ be
elements of $\mathcal{A}_{sa}$ and $\{p_k\}_{k=1}^n$ be a finite
subset of the set $\{p_\xi\}$ and $p=\sum_{k=1}^n p_k$. We have
the net of all finite sums of the kind $\sum_{k=1}^n p_k$ of
orthogonal projections of the set $\{p_\xi\}$ ultraweakly
converges to the identity operator in $B(H)$. Then for all $\xi$,
$\eta$ the element $\{p_\xi abp_\eta\}$ is an ultraweak limit in
$B(H)$ of the net $\{\sum_i p_\xi ap_i p_i bp_\eta\}$ of all
finite sums $\{\sum_{k=1}^n p_\xi ap_k p_k bp_\eta\}$ on all
subsets $\{p_k\}_{k=1}^n\subset \{p_\xi\}$, and the element
$\{p_\xi abp_\eta\}$ belongs to $\mathcal{A}$. Hence the assertion
of proposition 6 holds. $\triangleright$
\medskip
{\it Remark 2.} Let $A$ be a C$^*$-algebra on a Hilbert space $H$,
$\{p_i\}$ be an infinite orthogonal set of projections in $A$ with
the least upper bound $1$ in the algebra $B(H)$. Note that then by
lemma 4 in \cite{2} the order and the norm in the vector space
$\sum_{i,j}^\oplus p_i Ap_j$ can be introduced as follows: we
write $\{a_{ij}\}\geq 0$, if this element is zero or positive
element in $B(H)$ in the sense of the equality
$B(H)=\sum_{\xi,\eta}^\oplus q_\xi B(H)q_\eta$, where $\{q_\xi\}$
is an arbitrary maximal orthogonal set of minimal projections of
the algebra $B(H)$; $\Vert \{a_{ij}\}\Vert$ is equal to the norm
in $B(H)$ of this element in the sense of the equality
$B(H)=\sum_{\xi,\eta}^\oplus q_\xi B(H)q_\eta$ (example 1). By
lemmas 3 and 4 in \cite{2} they coincide with the order and the
norm defined in lemma 2 and proposition 3, correspondingly.
\medskip
{\it Remark 3.} Suppose that all conditions of remark 2 hold. Then
$B(H)\equiv\mathcal{B(H)}=\sum_{\xi,\eta}^\oplus q_\xi
B(H)q_\eta$, where $\mathcal{B(H)}=\{\{q_\xi aq_\eta\}: a\in
B(H)\}$. Also, we have $\sum_{ij}^\oplus p_i Ap_j$ is a Banach
space and an order unit space (lemma 2, Proposition 3). Suppose
that $\{q_\xi\}$ is such maximal orthogonal set of minimal
projections of the algebra $B(H)$ that $p_i=\sup_\eta q_\eta$, for
some subset $\{q_\eta\}\subset \{q_\xi\}$, for all $i$. Note that
$B(H)\equiv\{\{p_iap_j\}: a\in B(H)\}=\sum_{ij}^\oplus p_i
B(H)p_j$. By propositions 4 and 6 the order unit space
$\mathcal{A}=\{\{p_iap_j\}: a\in A\}$ is closed concerning the
associative multiplication of the algebra $\sum_{ij}^\oplus p_i
B(H)p_j$ (what is the same that $\mathcal{N}=\{\{p_iap_j\}: a\in
B(H)\}$).
At the same time, the order unit space $\sum_{ij}^\oplus p_i Ap_j$
is the order unit subspace of the algebra $\sum_{ij}^\oplus p_i
B(H)p_j$.
Since $B(H)\equiv\sum_{ij}^\oplus p_i B(H)p_j$, then
$\sum_{ij}^\oplus p_i B(H)p_j$ is a von Neumann algebra, and
without loss of generality, this algebra can be considered as the
algebra $B(H)$.
Note that if the space $\sum_{ij}^\oplus p_i Ap_j$ is closed
concerning the associative multiplication of the algebra
$\sum_{ij}^\oplus p_i B(H)p_j$, then $\sum_{ij}^\oplus p_i Ap_j$
is a C$^*$-algebra. Also, when we consider the C$^*$-algebra $A$
with the conditions which are listed above, then we have the
algebra $\sum_{ij}^\oplus p_i B(H)p_j$ (i.e. actually the algebra
$B(H)$) and the vector space $\sum_{ij}^\oplus p_i Ap_j$ as an
order unit subspace of the algebra $\sum_{ij}^\oplus p_i B(H)p_j$.
Then we have
$$
\mathcal{A}\subseteq \sum_{ij}^\oplus p_i Ap_j\subseteq
\sum_{ij}^\oplus p_i B(H)p_j.
$$
Thus, further, when we say that $\sum_{ij}^\oplus p_i Ap_j$ is a
C$^*$-algebra we assume that the vector space $\sum_{ij}^\oplus
p_i Ap_j$ is closed concerning the associative multiplication of
the algebra $\sum_{ij}^\oplus p_i B(H)p_j$.
The involution in the sense of the identification
$\sum_{ij}^\oplus p_i B(H)p_j\equiv B(H)$ coincides with the next
map:
$$
\{a_{ij}\}^*=\{a_{ji}^*\}, \{a_{ij}\}\in \sum_{ij}^\oplus p_i
B(H)p_j.
$$
Indeed, there exists an element $a\in B(H)$ such that
$a=\{a_{ij}\}=\{p_iap_j\}$. Then $a^*=\{p_ia^*p_j\}$ in the sense
of $B(H)\equiv \mathcal{N}$. We have $a_{ij}=p_iap_j$,
$a_{ij}^*=p_ja^*p_i$ for all $i$, $j$. Therefore
$\{p_ia^*p_j\}=\{a_{ji}^*\}$. Hence $a^*=\{a_{ji}^*\}$. Let
$(\sum_{ij}^\oplus p_i B(H)p_j)_{sa}=\{\{a_{ij}\}: \{a_{ij}\}\in
\sum_{ij}^\oplus p_i B(H)p_j, \{a_{ij}\}^*=\{a_{ij}\}\}$. Then
$$
\sum_{ij}^\oplus p_i B(H)p_j=(\sum_{ij}^\oplus p_i
B(H)p_j)_{sa}+i(\sum_{ij}^\oplus p_i B(H)p_j)_{sa}.
$$
\medskip
{\bf Lemma 7.} {\it Let $A$ be a C$^*$-algebra on a Hilbert space
$H$, $\{p_i\}$ be an infinite orthogonal set of projections of the
algebra $A$ with least upper bound $1$ in $B(H)$ and
$(\sum_{ij}^\oplus p_i Ap_j)_{sa}=\{\{a_{ij}\}: \{a_{ij}\}\in
\sum_{ij}^\oplus p_i Ap_j, \{a_{ij}\}^*=\{a_{ij}\}\}$. Then
$$
\sum_{ij}^\oplus p_i Ap_j=(\sum_{ij}^\oplus p_i
Ap_j)_{sa}+i(\sum_{ij}^\oplus p_i Ap_j)_{sa}. \,\,\,(**)
$$
In this case the equality $\{a_{ij}\}^*=\{a_{ij}\}$ holds for
$\{a_{ij}\}\in \sum_{ij}^\oplus p_i Ap_j$ if and only if
$a_{ij}^*=a_{ji}$ for all $i$, $j$.}
{\bf Proof.} Let $\{a_{ij}\}\in \sum_{ij}^\oplus p_i Ap_j$. We
have $a_{ij}+a_{ji}=a_1+ia_2$, where $a_1$, $a_2\in
(\sum_{ij}^\oplus p_i Ap_j)_{sa}$, for all $i$ and $j$, since
$a_{ij}+a_{ji}\in A$. Then
$a_{ij}+a_{ji}=p_ia_1p_j+p_ja_1p_i+i(p_ia_2p_j+p_ja_2p_i)$,
$a_1=p_ia_1p_j+p_ja_1p_i$, $a_2=p_ia_2p_j+p_ja_2p_i$ for all $i$
and $j$. Let $a^1_{ij}=p_ia_1p_j+p_ja_1p_i$,
$a^2_{ij}=p_ia_2p_j+p_ja_2p_i$ for all $i$ and $j$. Then by the
definition of the vector space $\sum_{ij}^\oplus p_i Ap_j$ we have
$\{a^1_{ij}\}$, $\{a^2_{ij}\}\in \sum_{ij}^\oplus p_i Ap_j$. In
this case $\{a^k_{ij}\}^*=\{a^k_{ij}\}$, $k=1,2$. Since the
element $\{a_{ij}\}\in \sum_{ij}^\oplus p_i Ap_j$ was chosen
arbitrarily we have the equality (**).
The rest part of the assertion of lemma 7 holds by the definition
of the self-adjoint elements $\{a^k_{ij}\}$, $k=1,2$.
$\triangleright$
\medskip
{\bf Lemma 8.} {\it Let $B(H)$ be a $*$-algebra of all bounded
linear operators on a Hilbert space $H$, $\{p_\xi\}$ be an
infinite orthogonal set of projections of $B(H)$ with the least
upper bound $1$. Then the associative multiplication of the
algebra $\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$ (i.e. of the
algebra $B(H)$) coincides with the associative multiplication
defined as follows:
$$
\cdot :<\{a_{\xi,\eta}\},\{b_{\xi,\eta}\}>\to \{\sum_i a_{\xi i}
b_{i\eta}\}, \{a_{\xi\eta}\},\{b_{\xi\eta} \}\in
(\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta).
$$
}
{\bf Proof.} Let $\{a_{\xi\eta}\},\{b_{\xi\eta} \}\in
(\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta)$. We have $B(H)\equiv
\mathcal{N}\equiv \sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$.
Therefore, it can be regarded that
$B(H)=\mathcal{N}=\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$. There
exists elements $a$, $b$ in the algebra $B(H)$ such that $p_\xi
ap_\eta=a_{\xi\eta}$, $p_\xi bp_\eta=b_{\xi\eta}$ for all $\xi$,
$\eta$. Therefore $\{a_{\xi\eta}\}=\{p_\xi ap_\eta\}$,
$\{b_{\xi\eta}\}=\{p_\xi bp_\eta\}$. Then by lemma 5 we have the
associative multiplication of the algebra $\sum_{\xi,\eta}^\oplus
p_\xi B(H)p_\eta$ (i.e. of the algebra $B(H)$) coincides with the
operation defined in lemma 8. $\triangleright$
\medskip
{\bf Proposition 9.} (\cite{1}){\it Let $A$ be a von Neumann
algebra on a Hilbert space $H$, $\{p_i\}$ be an infinite
orthogonal set of projections of the algebra $A$ with least upper
bound $1$ in $B(H)$. Then $A=\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$.}
{\bf Proof.}Let $a$ be an element of the vector space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ and $a=\{a_{\xi\eta}\}$,
where $a_{\xi\xi}=p_\xi ap_\xi$, $a_{\xi\eta}=p_\xi ap_\eta$ for
all $\xi$, $\eta$. We have $a\in B(H)=\sum_{\xi,\eta}^\oplus p_\xi
B(H)p_\eta$ and $(\sum_{k=1}^n p_k)a(\sum_{k=1}^n p_k)\in A$ for
any $\{p_k\}_{k=1}^n\subset \{p_\xi\}$. Let
$$
b_n^\alpha=\sum_{kl=1}^n p_k^\alpha ap_l^\alpha=(\sum_{kl=1}^n
p_k^\alpha)a(\sum_{kl=1}^n p_k^\alpha)
$$ for all natural numbers $n$ and finite subsets $\{p_k^\alpha\}_{k=1}^n\subset
\{p_i\}$. Then by the proof of lemma 3 in \cite{2} the net
$(b_n^\alpha)$ ultraweakly converges to $a$ in $B(H)$. At the same
time $A$ is ultraweakly closed in $B(H)$. Therefore $a\in A$ and
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta\subseteq A$.
$\triangleright$
{\bf Lemma 10.} {\it Let $A$ be a C$^*$-algebra on a Hilbert space
$H$, $\{p_\xi\}$ be an infinite orthogonal set of projections in
$A$ with the least upper bound $1$ in the algebra $B(H)$. Then, if
projections of the set $\{p_\xi\}$ are pairwise equivalent and for
every index $\xi$ the component $p_\xi Ap_\xi$ is a von Neumann
algebra, then the vector space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is closed concerning the multiplication of the algebra
$\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$ and
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a C$^*$-algebra.}
{\bf Proof.} First, note that $(p_\xi+p_\eta)A(p_\xi+p_\eta)$ is a
von Neumann algebra. Indeed, for any net $(a_\alpha)$ in $p_\xi
Ap_\eta$, weakly converging in $B(H)$ the net $(a_\alpha
x_{\xi\eta}^*)$ belongs to $p_\xi Ap_\xi$, where $x_{\xi\eta}$ is
an isometry in $A$ such that $x_{\xi\eta}x_{\xi\eta}^*=p_\xi$,
$x_{\xi\eta}^*x_{\xi\eta}=p_\eta$. Then since the net $(a_\alpha
x_{\xi\eta}^*)$ weakly converges in $B(H)$ then the weak limit $b$
in $B(H)$ of the net $(a_\alpha x_{\xi\eta}^*)$ belongs to $p_\xi
Ap_\xi$. Hence $bx_{\xi\eta}\in p_\xi Ap_\eta$. It is easy to see
that $bx_{\xi\eta}$ is a weak limit in $B(H)$ of the net
$(a_\alpha)$. Hence $p_\xi Ap_\eta$ is weakly closed in $B(H)$.
Let $\{a_{\xi\eta}\}$, $\{b_{\xi\eta}\}\in (\sum_{\xi\eta}^\oplus
p_\xi Ap_\eta)$. We have
$$
\sum_{\xi\eta}^\oplus p_\xi Ap_\eta\subseteq \sum_{\xi\eta}^\oplus
p_\xi B(H)p_\eta=B(H).
$$
Therefore there exist elements $a$, $b$ in the algebra
$\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$ (i.e. in the algebra
$B(H)$) such that $p_\xi ap_\eta=a_{\xi\eta}$, $p_\xi
bp_\eta=b_{\xi\eta}$ for all $\xi$, $\eta$. Therefore
$\{a_{\xi\eta}\}=\{p_\xi ap_\eta\}$, $\{b_{\xi\eta}\}=\{p_\xi
bp_\eta\}$. We have
$$
\sum_i a_{\xi i} b_{i\eta}=p_\xi abp_\eta
$$
calculated in $\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$ belong to
$p_\xi Ap_\eta$. Since the indexes $\xi$, $\eta$ were chosen
arbitrarily and the product $\{p_\xi ap_\eta\}\{p_\xi
bp_\eta\}=ab$ belongs to $\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$,
then the product of the elements $a$ and $b$ belongs to
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. Therefore the vector space
$\sum_{\xi\eta}^\oplus p_\xi Ap_\eta$ is closed with respect to
the associative multiplication of the algebra
$\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$. At the same time,
$\sum_{\xi\eta}^\oplus p_\xi Ap_\eta$ is a norm closed subspace of
the algebra $\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta=B(H)$. Hence
$\sum_{\xi\eta}^\oplus p_\xi Ap_\eta$ is a C$^*$-algebra and the
multiplication in $\sum_{\xi\eta}^\oplus p_\xi Ap_\eta$ can be
defined as in the formulation of lemma 8. $\triangleright$
\medskip
{\bf Theorem 11.} {\it Let $A$ be a C$^*$-algebra on a Hilbert
space $H$, $\{p_\xi\}$ be an infinite orthogonal set of
projections in $A$ with the least upper bound $1$ in the algebra
$B(H)$. Suppose that projections of the set $\{p_\xi\}$ are
pearwise equivalent and for any $\xi$ $p_\xi Ap_\xi$ is a von
Nemann algebra. Then $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a
von Neumann algebra. }
{\bf Proof.} Let $\{x_{\xi\eta}\}$ be such set of isometries in
$A$ that $p_\xi=x_{\xi\eta}x_{\xi\eta}^*$,
$p_\eta=x_{\xi\eta}^*x_{\xi\eta}$ for all $\xi$, $\eta$. Let
$\xi$, $\eta$ be arbitrary indexes. We prove that $p_\xi Ap_\eta$
is weakly closed. We have $p_\xi Ap_\eta p_\eta Ap_\xi \subseteq
p_\xi Ap_\xi$ and $p_\xi Ap_\eta =x_{\xi\eta}Ax_{\xi\eta}$. Let
$(a_\alpha)$ be a net in $p_\xi Ap_\eta$, weakly converging to an
element $a$ in $B(H)$. Then the exists a net $(b_\alpha)$ in
$p_\xi Ap_\eta$ such that $a_\alpha=x_{\xi\eta}b_\alpha
x_{\xi\eta}$ for all $\alpha$. By the weakly continuity of the
multiplication separately on multipliers the net $(a_\alpha
x_{\xi\eta}^*)$ weakly converges to the element $ax_{\xi\eta}$ in
the algebra $B(H)$. Since $(a_\alpha x_{\xi\eta}*)\subset p_\xi
Ap_\xi$ and $p_\xi Ap_\xi$ is weakly closed in $B(H)$, then
$ax_{\xi\eta}^*\in p_\xi Ap_\xi$. Hence there exists an element
$b\in A$ such that
$ax_{\xi\eta}^*=x_{\xi\eta}bx_{\xi\eta}x_{\xi\eta}^*$. Then
$ax_{\xi\eta}^*x_{\xi\eta}=x_{\xi\eta}bx_{\xi\eta}x_{\xi\eta}^*x_{\xi\eta}
=x_{\xi\eta}bx_{\xi\eta}p_\eta=x_{\xi\eta}bx_{\xi\eta}\in p_\xi
Ap_\eta$. At the same time $a_\alpha p_\eta=a_\alpha$ for all
$\alpha$. Hence, $ap_\eta=a$ in the algebra $B(H)$. Since
$a=ax_{\xi\eta}^*x_{\xi\eta}=x_{\xi\eta}bx_{\xi\eta}\in p_\xi
Ap_\eta$, then $a\in p_\xi Ap_\eta$. Since the net $(a_\alpha)$ is
chosen arbitrarily, then the component $p_\xi Ap_\eta$ is weakly
closed in $B(H)$. Let $(a_\alpha)$ be a net in
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, weakly converging to an
element $a$ in $B(H)$. Then for all $\xi$ and $\eta$ the net
$(p_\xi a_\alpha p_\eta)$ weakly converges to $p_\xi ap_\eta$ in
$B(H)$. In this case, by the previous part of the proof $p_\xi
ap_\eta\in p_\xi Ap_\eta$ for all $\xi$, $\eta$. Note that $a\in
\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$. Hence $a\in
\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. Since the net $(a_\alpha)$
is chosen arbitrarily, then the vector space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is weakly closed in the
algebra $\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta\equiv B(H)$.
Therefore by lemma 10 $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a
von Neumann algebra. $\triangleright$
\medskip
{\bf Proposition 12.} {\it Let $A$ be a monotone complete
C$^*$-algebra on a Hilbert space $H$, $\{p_\xi\}$ be an infinite
orthogonal set of projections in $A$ with the least upper bound
$1$ in the algebra $B(H)$. Then the order unit space
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is monotone complete.}
{\bf Proof.} We have the C$^*$-subalgebra $p_\xi Ap_\xi$ is
monotone complete for any index $\xi$. Let $\{p_k\}_{k=1}^n$ be a
finite subset of the set $\{p_\xi\}$ and $p=\sum_{k=1}^n p_k$.
Then the C$^*$-subalgebra $pAp$ is also monotone complete.
Let $(a_\alpha)$ be a bounded monotone increasing net in
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. Since for any finite
subset $\{p_k\}_{k=1}^n\subseteq \{p_\xi\}$ the subalgebra
$(\sum_{k=1}^n p_k) A(\sum_{k=1}^n p_k)$ is monotone complete then
$$
\sup_\alpha (\sum_{k=1}^n p_k)a_\alpha (\sum_{k=1}^n p_k)\in
(\sum_{k=1}^n p_k) A(\sum_{k=1}^n p_k).
$$
Hence, $\{a_{\xi\eta}\}=\{\sup_\alpha p_\xi a_\alpha p_\xi\}\cup
\{p_\xi(\sup_\alpha (p_\xi+p_\eta) a_\alpha
(p_\xi+p_\eta))p_\eta\}_{\xi\neq\eta}$ is an element of the order
unit space $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. It can be
checked straightforwardly using the definition of the order in the
order unit space $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ that the
element $\{a_{\xi\eta}\}$ is the least upper bound of the net
$(a_\alpha)$. Since the net $(a_\alpha)$ was chosen arbitrarily
then the order unit space $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$
is monotone compete. $\triangleright$
\medskip
{\bf Theorem 13.} {\it Let $A$ be a monotone complete
C$^*$-algebra of bounded linear operators on a Hilbert space $H$,
$\{p_\xi\}$ be an infinite orthogonal set of projections in $A$
with the least upper bound $1$ in the algebra $B(H)$. Suppose that
projections of the set $\{p_\xi\}$ are pearwise equivalent and $A$
is not a von Nemann algebra. Then $A\neq \sum_{\xi,\eta}^\oplus
p_\xi Ap_\eta$ (i.e. $\mathcal{A}:=\{\{p_\xi ap_\eta\}: a\in
A\}\neq \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$). }
{\bf Proof.} We have there exists a bounded monotone increasing
net $(a_\alpha)$ of elements in $A$, the least upper bound
$\sup_{A} a_\alpha$ in the algebra $A$ and the least upper bound
$\sup_{\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta} a_\alpha$ in the
algebra $\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$ of which are
different. Otherwise $A$ is a von Nemann algebra.
By the definition of the order in the algebra
$\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$ there exists a projection
$p\in \{p_\xi\}$ such that the least upper bound $\sup_{pAp}
pa_\alpha p$ in the algebra $pAp$ and the least upper bound
$\sup_{pB(H)p} p a_\alpha p$ in the algebra $pB(H)p$ of the
bounded monotone increasing net $(pa_\alpha p)$ of elements in
$pAp$ are different. Indeed, let $a=\sup_A a_\alpha$,
$b=\sup_{\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta} a_\alpha$. Since
$A\subseteq \sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$, then $b\leq
a$ and $0\leq a-b$. Hence, if $p_\xi (a-b) p_\xi=0$ for all $\xi$,
then $p_\xi (a-b)=(a-b)p_\xi=0$. Therefore by lemma 2 in \cite{2}
$a-b=0$, i.e. $a=b$. Hence $pAp$ is not a von Nemann algebra.
We have there exists an infinite orthogonal set $\{e_i\}$ of
projections in $pAp$, the least upper bound $\sup_{pAp} e_i$ in
the algebra $pAp$ and the least upper bound $\sup_{pB(H)p} e_i$ in
the algebra $pB(H)p$ of which are different. Otherwise $pAp$ is a
von Neumann algebra.
Indeed, any maximal commutative subalgebra $A_o$ of $pAp$ is
monotone complete. For any normal positive linear functional
$\rho\in B(H)$ and for any infinite orthogonal set $\{q_i\}$ of
projections in $A_o$ we have $\rho(\sup_i q_i)=\sum_i \rho(q_i)$,
where $\sup_i q_i$ is the least upper bound of the set $\{q_i\}$
in $A_o$. Hence by the theorem on extension of a $\sigma$-additive
measure to a normal linear functional $\rho\vert_{A_o}$ is a
normal functional on $A_o$. Hence $A_o$ is a commutative von
Neumann algebra. At the same time the maximal commutative
subalgebra $A_o$ of the algebra $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is chosen arbitrarily. Therefore by \cite{3}
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ is a von Neumann algebra.
What is impossible.
Let $\{x_{\xi\eta}\}$ be such set of isometries in $A$ that
$p_\xi=x_{\xi\eta}x_{\xi\eta}^*$,
$p_\eta=x_{\xi\eta}^*x_{\xi\eta}$ for all $\xi$, $\eta$, and let
$p_1=p$. Let $\{x_{1\xi}\}$ be the subset of the set
$\{x_{\xi\eta}\}$ such that $p_1=x_{1\xi}x_{1\xi}^*$,
$p_\xi=x_{1\xi}^*x_{1\xi}$ for all $\xi$. Without loss of
generality we regard that the set of indexes $i$ for $\{e_i\}$ is
a subset of the set of indexes $\xi$ for $\{p_\xi\}$. Let
$\{e_ix_{1i}\}$ be a set of all components of some infinite
dimensional matrix $\{a_{\xi\eta}\}$, the components, which are
does not present, are zeros and $\{x_{1i}^*e_i^*\}$ be also an
analogous matrix, which coincides with $\{a_{\xi\eta}^*\}$. We
have $\sum_i e_ix_{1i}x_{1i}^*e_i^*=\sum_i e_ip_1e_i^*=\sum_i
e_ie_i^*=\sum_i e_i\leq \sup_{pAp} e_i$. Therefore
$\{a_{\xi\eta}\}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. Then
$\{a_{\xi\eta}^*\}\in \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$.
Therefore if $\{a_{\xi\eta}\}\in A$ (i.e. in
$\mathcal{A}:=\{\{p_\xi ap_\eta\}: a\in A\}$) then the product
$\{a_{\xi\eta}\}\cdot\{a_{\xi\eta}^*\}$ in $\sum_{ij}^\oplus p_i
B(H)p_j$ belongs to $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$. In
this case we have the infinite dimensional matrix
$\{a_{\xi\eta}\}\cdot\{a_{\xi\eta}^*\}$ contains the component
$\sum_i e_ix_{1i}\cdot x_{1i}^*e_i^*$ such that $\sum_i
e_ix_{1i}\cdot x_{1i}^*e_i^*=p_1(\sum_i e_ix_{1i}\cdot
x_{1i}^*e_i^*)p_1$. Consequently,
$p_1(\{a_{\xi\eta}\}\cdot\{a_{\xi\eta}^*\})p_1=\sum_i
e_ix_{1i}\cdot x_{1i}^*e_i^*$. Hence $\sum_i e_ix_{1i}\cdot
x_{1i}^*e_i^*\in p_1 (\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta)p_1=p_1Ap_1$. Since $\sum_i e_ix_{1i}\cdot
x_{1i}^*e_i^*=\sum_i e_ip_1e_i^*=\sum_i e_ie_i^*=\sum_i e_i$, then
$\sum_i e_i\in p_1Ap_1$, i.e. $\sup_{pB(H)p} e_i\in p_1Ap_1$. The
last statement is a contradiction. Therefore
$\{a_{\xi\eta}\}\notin A$. Hence $A\neq \sum_{\xi,\eta}^\oplus
p_\xi Ap_\eta$ (i.e. $\mathcal{A}:=\{\{p_\xi ap_\eta\}: a\in
A\}\neq \sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$). $\triangleright$
\medskip
The next assertion follows by theorem 13 and it's proof.
{\bf Corollary 14.} {\it Let $A$ be a C$^*$-algebra on a Hilbert
space $H$, $\{p_\xi\}$ be an infinite orthogonal set of
projections in $A$ with the least upper bound $1$ in the algebra
$B(H)$. Suppose that the order unit space $\sum_{\xi,\eta}^\oplus
p_\xi Ap_\eta$ is monotone complete and there exists a bounded
monotone increasing net $(a_\alpha)$ of elements in
$\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$, the least upper bound
$\sup_{\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta} a_\alpha$ in the
algebra $\sum_{\xi,\eta}^\oplus p_\xi Ap_\eta$ and the least upper
bound $\sup_{\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta} a_\alpha$ in
the algebra $\sum_{\xi\eta}^\oplus p_\xi B(H)p_\eta$ of which are
different. Then the vector space $\sum_{\xi,\eta}^\oplus p_\xi
Ap_\eta$ is not closed concerning the multiplication of the
algebra $\sum_{\xi,\eta}^\oplus p_\xi B(H)p_\eta$.}
\bigskip
\section{Application}
\medskip
Let $n$ be an infinite cardinal number, $\Xi$ a set of indexes of
cardinality $n$. Let $\{e_{ij}\}$ be a set of matrix units such
that $e_{ij}$ is a $n\times n$-dimensional matrix, i.e.
$e_{ij}=(a_{\alpha\beta})_{\alpha\beta\i\Xi}$, whose $(i,j)$-s
component is $1$, i.e. $a_{ij}=1$, and the rest components are
zero. Let $X$ be a huperstonean compact, $C(X)$ the commutative
algebra of all complex-valued continuous functions on the compact
$X$ and
$$
\mathcal{M}=\{\sum_{ij\in\Xi}\lambda_{ij}(x)e_{ij}: (\forall ij
\lambda_{ij}(x)\in C(X))
$$
$$
(\exists K\in R)(\forall m\in N)(\forall
\{e_{kl}\}_{kl=1}^m\subseteq \{e_{ij}\})\Vert\sum_{kl=1\dots
m}\lambda_{kl}(x)e_{kl}\Vert\leq K\},
$$
where $\Vert\sum_{kl=1\dots m}\lambda_{kl}(x)e_{kl}\Vert\leq K$
means $(\forall x_o\in X) \Vert \sum_{kl=1\dots
m}\lambda_{kl}(x_o)e_{kl}\Vert\leq K$. The set $\mathcal{M}$ is a
vector space with pointwise algebraic operations. The map
$\Vert\,\,\, \Vert : \mathcal{M}\to {\bf R}_+$ defined as
$$
\Vert a \Vert = \sup_{\{e_{kl}\}_{kl=1}^n\subseteq
\{e_{ij}\}}\Vert\sum_{kl=1}^n \lambda_{kl}(x)e_{kl}\Vert,
$$
is a norm on the vector space $\mathcal{M}$, where $a\in
\mathcal{M}$ and $a=\sum_{ij\in\Xi}\lambda_{ij}(x)e_{ij}$.
{\bf Theorem 15.} {\it $\mathcal{M}$ is a von Neumann algebra of
type I$_n$ and $\mathcal{M}=C(X)\otimes M_n({\bf C})$.}
{\it Proof.} It is easy to see that the set $\mathcal{M}$ is a
vector space with the componentwise algebraic operations. It is
known that the vector space $C(X,M_n({\bf C}))$ of continuous
matrix-valued maps on the compact $X$ is a C$^*$-algebra. Let
$A=C(X,M_n({\bf C}))$ and $e_i$ be a constant $e_{ii}$-valued map
on $X$, i.e. $e_i$ is a projection of the algebra $A$. Then
$\{e_i\}$ is an orthogonal set of projections with $\sup_i e_i=1$
in the algebra $A$. Then $\sum_{ij}^\oplus e_iAe_j=\mathcal{M}$.
We have a C$^*$-algebra $A$ can be embedded in $B(H)$ for some
Hilbert space $H$. Then $\sum_{ij}^\oplus e_iAe_j$ can be
embedded in $B(H)$. For any $i$ $e_iAe_i=C(X)e_i$, i.e. the
component $e_iAe_i$ is weakly closed in $B(H)$. Hence, by theorem
11 the image of vector space $\mathcal{M}$ in $B(H)$ is a von
Neumann algebra. Hence, $\mathcal{M}$ is a von Neumann algebra.
Note, that the set $\{e_i\}$ is a maximal orthogonal set of
abelian projections with central support 1. Hence, $\mathcal{M}$
is a von Neumann algebra of type I$_n$. Moreover the center
$Z(\mathcal{M})$ of the algebra $\mathcal{M}$ is isomorphic to
$C(X)$ and $\mathcal{M}=C(X)\otimes M_n({\bf C})$.
$\triangleright$
\bigskip
|
\section{Introduction}
This paper uses methods from Conley index theory to establish a continuation
of isolated attractors. For these attractors there is a stable neighborhood,
i.e. a positive invariant isolating neighborhood. Traditionally the
Conley index is applied to isolated invariant set whose isolating
neighborhoods are bounded and satisfy some compactness assumption,
Rybakowski \cite{Rybakowski1987} calls them admissible neighborhoods.
Focusing on a stable neighborhood we show that a non-autonomously
perturbed system has also a stable neighborhood which is {}``close''
to the original one.
A similar result for parabolic PDEs was obtained by Prizzi in \cite{Prizzi2005}
for small almost-periodic perturbations with sufficiently high {}``frequency''.
Because almost-periodicity causes compactness of the {}``perturbation-space''
he can obtain non-empty invariant sets even for the unstable case
(for flows on locally compact spaces our result also applies to totally
unstable invariant sets). In \cite{Ward1994} Ward obtains the result
for ODEs and small perturbations satisfying some hypothesis (H1).
Our result generalizes both of them in a way that we don't need almost-periodicity
and applies to quite general semiflows even in the infinite-dimensional
setting.
An open question to us is if an unstable invariant set with non-trivial
Conley index disappears for all small non-autonomous perturbations.
The continuation result and the Wa{\.z}ewski principle only give
us a non-empty positive invariant set. Obviously this cannot happen
in dimension $1$ and the perturbation should neither be almost-periodic
nor satisfy Ward's hypothesis (H1).
Furthermore, our result implies the upper semicontinuity of the global
pullback attractor (see \cite{Caraballo2003}) if the perturbation
is {}``uniformly small''. In addition, this also holds for local
attractors (called past attractors in \cite{Rasmussen2007}).
The result (see section \ref{sec:Inf}) is a consequence of the translation
invariance of the unperturbed flow and standard continuation results
of the Conley index. The proof is essentially contained in {\cite[Theorem 12.3]{Rybakowski1987}}
after replacing all admissibility arguments by an appropriate version
(see also \cite{Carbinatto2002}).
For locally compact metric spaces $X$ we can use ideas from \cite{Benci1991}.
Because the proof is very clear and easy to understand, we are going
to show it in section \ref{sec:Fin} even though this case, which
we call finite-dimensional case, is contained in the infinite-dimensional
case. Another reason to give the proof is that the obtained stable
neighborhoods are flow-defined and thus applicable to random dynamical
system on locally compact metric spaces without further assumptions
(see section \ref{sub:rds}). This implies that sufficiently small
bounded noise does not destroy a local attractor. We don't need regularity
of the support of the noise used by Ruelle in \cite{Ruelle1981} and
thus we could generalize his result (see also {\cite[Appendix D]{Bonatti2005}}
and the reference therein).
Furthermore, using the ideas of \cite{Kell2011c} and adjusting the
definition of semi-singular admissibility the infinite-dimensional
case can be extended to the discrete time setting. Thus stable neighborhoods
of a local attractor of a discrete time dynamical system, i.e. a continuous
maps $f:X\to X$, can be continued under small non-autonomous perturbations
$\tilde{f}:\mathbb{Z}\times X\to\mathbb{Z}\times X$.
\subsection*{Preliminaries}
\begin{defn}
[local semiflow] Let $\pi:D\to X$ be a continuous map into a topological
space $X$ and $D$ open in $\mathbb{R}^{+}\times X$ with $\{0\}\times X\subset D$.
For all $x\in X$ we define \[
\omega_{x}=\sup\{t>0\,|\,(t,x)\in D\}\in(0,\infty].\]
Assume $D\cap\mathbb{R}\times\{x\}=[0,\omega_{x})\times\{x\}$. Then
$\pi$ is called a (local) semiflow if the following holds
\begin{itemize}
\item $x\pi0=x$ for all $x\in X$
\item if $(t,x)\in D$ and $(s,x\pi t)\in D$ then $x\pi(s+t)$ is defined
and equals $(x\pi t)\pi s$
\end{itemize}
If, in addition, $\omega_{x}=\infty$ for all $x\in X$ then $\pi$
is called a global semiflow.\end{defn}
\begin{rem*}
We use the notation $x\pi t$ for $\pi(t,p)$ and $x\phi^{p}t=\phi(t,p,x)$
whenever $(t,p,x)\in D$ (see below). Furthermore, we write $x\phi^{p}[0,t_{0}]$
for the set $\{x\phi^{p}t\,|\, t\in[0,t_{0}]\}$ under the condition
that $t_{0}<\omega_{x}^{p}$, otherwise $x\phi^{p}[0,t_{0}]$ is not
defined.\end{rem*}
\begin{defn}
[non-autonomous dynamical system (NDS)] Let $X$ be a metric space
and $P$ be a set called the base set. A (local) NDS is a pair of
mappings\begin{eqnarray*}
\theta:\mathbb{R}\times P & \to & P\\
\phi:D & \to & X\end{eqnarray*}
such that the following holds:
\begin{itemize}
\item $\theta$ is a (not necessarily continuous) flow, called the base
flow
\item $D$ is open in $\mathbb{R}^{\ge0}\times P\times X$ with $\{0\}\times P\times X\subset D$,
$x\phi^{p}0=x$ and whenever $(s,p,x)\in D$ and $(s+t,p,x)\in D$
for some $s,t\ge0$\[
(t,p\theta s,x\phi^{p}s)\in D\]
and \[
x\phi^{p}(t+s)=(x\phi^{p}s)\phi^{p\theta s}t.\]
Furthermore, we define \[
\omega_{x}^{p}=\sup\{t>0\,|\,(t,p,x)\in D\ \}\]
\item $\phi$ is continuous with respect to $t\in\mathbb{R}_{\ge0}$ and
$x\in X$ for fixed $p\in P$.
\end{itemize}
We call $(\phi,\theta)$ a (local) semiprocess if $\theta$ is continuous
and $P$ a metric space.\end{defn}
\begin{rem*}
To every (semi)process $(\phi,\theta)$ we can associated a semiflow
$\pi:D\to P\times X$ in the following way. If $(t,p,x)\in D$ then
$(p,x)\pi t=(p\theta t,x\phi^{p}t)$. The resulting (semi)flow is
called skew product semiflow.
\end{rem*}
\begin{rem*}
If $P$ is a compact metric space and $\theta$ continuous then standard
Conley index theory is applicable: The unperturbed system $\pi$ is
a product of the flow $\theta$ and a (semi)flow $\pi_{0}$ on $X$
so that a Conley index $h(N,\pi_{0})$ lifts to \[
h(P\times N,\theta\times\pi_{0})=h(P,\theta)\wedge h(N,\pi_{0}).\]
The stability result follows from a standard continuation result for
the index and $H_{0}(h(P\times N,\theta\times\pi_{0}))\ne0$ iff $H_{0}(h(N,\pi_{0}))\ne0$
(see \cite{Kell2010} for classification of stability via the zeroth
singular homology $H_{0}$ of the index).
\end{rem*}
Although most of the following can be formulated for general NDS if
we assume uniform convergence w.r.t. to the orbit set of $\theta$
(see section \ref{sub:non-auto}) we will restrict our attention to
$P=\mathbb{R}$ and $\tau_{t}(s):=\theta(t,s)=t+s$. This includes
processes generated by non-autonomous differential equations. Furthermore,
we will only look at the induced skew product (semi)flow, i.e. $\pi:D\subset\mathbb{R}^{\ge0}\times\tilde{X}\to\tilde{X}$
with $\tilde{X}=\mathbb{R}\times X$.
A map $\sigma:J\to X$ with $J\subset\mathbb{R}$ is called a solution
of $\pi$ through $x_{0}\in X$ if $0\in J$ and $\sigma(0)=x_{0}$
and whenever $t,t+s\in J$ for some $s>0$ with $s<\omega_{\sigma(t)}$
then \[
\sigma(t+s)=\sigma(t)\pi s.\]
$\sigma$ is a left solution if $J\cap\mathbb{R}^{-}=(a,0]$ for some
$a\in[-\infty,0)$ and it is called a full left solution if $\mathbb{R}^{-}\subset J$.
Let $Y\subset X$ be arbitrary. We define the following sets\begin{align*}
A^{+}(Y) & =\{x\in Y\,|\, x\pi t\in Y\,\text{for all }t\in[0,\omega_{x})\}\\
A^{-}(Y) & =\{x\in Y\,|\,\text{\ensuremath{\exists}a full left solution \ensuremath{\sigma\,}through \ensuremath{x\,}with \ensuremath{\sigma(\mathbb{R}^{-})\subset Y}}\}\end{align*}
and \[
A(Y)=A^{-}(Y)\cap A^{+}(Y).\]
$Y$ is called invariant if $Y=A(Y)$, positive invariant if $Y=A^{+}(Y)$
and negative invariant if $Y=A^{-}(Y)$. The sets $A(Y)$ and $A^{\pm}(Y)$
depend on $Y$ and $\pi$. In case we talk about several flows $\pi_{n}$
we will write $A_{\pi_{n}}(Y)$ and $A_{\pi_{n}}^{\pm}(Y)$.
If $S\subset\operatorname{int}N$ for some closed neighborhood $N$
and $S$ is the maximal invariant set in $N$, i.e. $S=A(N)$, then
$S$ is called an isolated invariant set (w.r.t. $\pi$). A closed
set $N$ is called isolating neighborhood if the maximal $\pi$-invariant
set is in the interior of $N$, i.e. $A(N)\subset\operatorname{int}N$.
In particular the closure $\operatorname{cl}U$ of a neighborhood
$U$ of $A(N)$ with $U\subset N$ is an isolating neighborhood.
\begin{defn}
A closed isolating neighborhood $N$ is called stable if \[
A(N)\subset\operatorname{int}N\]
and $N$ is positive invariant, i.e. \[
A^{+}(N)=N.\]
\end{defn}
\section{\label{sec:Fin}Finite dimensional case}
Although many ideas in this section are taken from Conley index theory
we don't want to introduce the full theory of the Conley index. Many
of the techniques and definitions in this section are based on Benci's
paper \cite{Benci1991}. From now on let $X$ be locally compact.
Since our isolating neighborhood will be compact we will assume that
$\pi_{0}$ is a flow in order to simplify our arguments, i.e. there
will be no finite-time blow-up. This will be true for the skew product
flow induced by a process if the {}``$X$-component'' of set $\tilde{N}\subset\mathbb{R}\times X$
is compact as well (the unbounded component represents time).
From \cite{Benci1991} we take the following definitions:\[
G^{T}(N)=G_{\pi}^{T}(N)=\{x\in X\,|\, x\pi[-T,T]\subset\operatorname{cl}N\}\]
and\[
\Gamma^{T}(N)=\Gamma_{\pi}^{T}(N)=\{x\in G^{T}(N)\,|\, x\pi[0,T]\cap\partial N\ne\varnothing\}.\]
Furthermore we define the set of isolating neighborhoods as \[
\mathcal{F}=\mathcal{F}_{\pi}=\{N\subset X\,|\,\operatorname{int}N\ne\varnothing\,\mbox{and}\,\exists T>0\,\mbox{s.t.}\, G^{T}(N)\subset\operatorname{int}N\}.\]
\begin{lem}
[\cite{Benci1991}]If $N\in\mathcal{F}$ then the following hold:
\begin{enumerate}
\item $T_{1}>T_{2}>0$ then $G^{T_{1}}(N)\subset G^{T_{2}}(N)$
\item $G^{T}(N)$ and $\Gamma^{T}(N)$ are closed and $\Gamma^{T}(N)\subset\partial G^{T}(N)$
\item If $G^{T}(N)\subset\operatorname{int}N$ then $G^{2T}(N)\subset\operatorname{int}G^{T}(N)$
\item If $\Gamma^{T}(N)=\varnothing$ then $A^{+}(G^{T}(N))=G^{T}(N)$
\end{enumerate}
\end{lem}
\begin{rem*}
It can be shown that for large $T$ the pair $(G^{T}(N),\Gamma^{T}(N))$
defines an index pair (see section \ref{sec:Inf} for the definition).
We will focus only on stable neighborhoods for which $\Gamma^{T}(N)=\varnothing$
for large $T$. We can show that for arbitrary isolated invariant
set with non-trivial Conley index the index continues even for sufficiently
small non-autonomous perturbations. But the corresponding index pair
$(N_{1},N_{2})$ is unbounded. In particular, we cannot show in general
that $A(N_{1})\ne0$ if the index $(N_{1}/N_{2},[N_{2}])$ is non-trivial.
\end{rem*}
A standard result from Conley index theory is the following theorem.
This result also holds under the assumption that $N$ is strongly
$\pi$-admissible (see section \ref{sec:Inf} for the definition).
\begin{thm}
[{\cite[Corollary 5.5]{Rybakowski1987}}]\label{thm:stable} Let $N$
be a compact isolating neighborhood such that \[
A^{-}(N)=A(N)=K\ne\varnothing\]
then there exist a stable isolating neighborhood $B\subset\operatorname{int}N$
such that for all $x\in\partial B$ there is an $\epsilon>0$ such
that for all $t\in(0,\epsilon)$\[
x\pi(-t)\notin B\]
and \[
x\pi t\in\operatorname{int}B.\]
An isolating neighborhood with this property will be called stable
isolating block. \end{thm}
\begin{rem*}
Suppose we got $B'\subset N$ as a result from the theorem. Applying
the theorem again we get a second stable isolating block $B\subset B'$.
Because $B'$ is compact and $K$ is in its interior, $U_{\delta}(B)=\{x\in X\,|\, d(x,B)<\delta\}\subset B'$
for some $\delta>0$.
\end{rem*}
Let $\pi_{0}$ be a flow on $X$. Then we can define a skew product
flow $\pi$ on $\mathbb{R}\times X$ by \[
(s,x)\pi t=(\tau_{s}(t),x\pi_{0}t).\]
Thus $\pi_{0}$ can be considered as a process which does not depend
on the base flow $\tau$ and is therefore translation invariant w.r.t.
time.
\begin{defn}
[Semi-singular convergence]Let $\pi_{n}$ be a sequence of semiflows
on $\mathbb{R}\times X$ and $(s_{n},x_{n})_{n\in\mathbb{N}}$ be
any sequence in $\mathbb{R}\times X$ and $t_{n}\in\mathbb{R}^{\ge0}$
such that $x_{n}\to x_{0}$ and $t_{n}\to t_{0}$ then we say $\pi_{n}$
converges semi-singularly to $\pi_{0}$ if \[
P_{2}((s_{n},x_{n})\pi_{n}t_{n})\to x_{0}\pi_{0}t_{0}\]
where $P_{2}(s,x)=x$. We write $\pi_{n}\overset{\mbox{\tiny ssing}}{\longrightarrow}\pi_{0}$.
Semi-singular convergence implies that $\pi_{n}\to\pi$ in the usual
sense if each $\pi_{n}$ is skew product flow of processes because
if in addition $s_{n}\to s_{0}$ for some $s_{0}\in\mathbb{R}$ then
\[
(s_{n},x_{n})\pi_{n}t_{n}=(s_{n}+t_{n},P_{2}((s_{n},x_{n})\pi_{n}t_{n}))\to(s_{0}+t_{0},x_{0}\pi_{0}t_{0})=(s_{0},x_{0})\pi t_{0}.\]
\end{defn}
\begin{rem*}
Semi-singular convergence is a simplified version of singular convergence
defined in \cite{Carbinatto2002}. With the metric $d_{\epsilon}(t,s)=\epsilon\min\{|t-s|,1\}$
we recover the singular convergence.
\end{rem*}
Assume from now on that $\pi_{n}$ is a sequence of flows such that
$\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$.
\begin{example*}
The standard example is an autonomous ordinary differential equation
\[
\dot{x}=f_{0}(x)\]
and the non-autonomous ODEs\[
\dot{x}=f_{n}(t,x)\]
such that $\sup_{t}\|f_{n}(t,x)-f_{0}(x)\|<\epsilon_{n}$ with $\epsilon_{n}\to0$.
If $f_{0}$ and $f_{n}$ are locally Lipschitz continuous then they
induce a local flow $\pi_{0}$, resp. local processes $\phi_{n}$.
Restricted to some compact set in $X$ we can assume that $\pi_{0}$
and $\phi_{n}$ are defined everywhere. Suppose we have $(s_{n},x_{n})_{n\in\mathbb{N}}$
with $x_{n}\to x_{0}$ and $t_{n}\to t_{0}$. The ODEs defined by\[
\dot{x}=\tilde{f}_{n}(t,x)=f_{n}(s_{n}+t,x)\]
generate time-translated processes $\tilde{\phi}_{n}$. We still have
$\sup_{t}\|\tilde{f}_{n}(t,x)-f_{0}(x)\|<\epsilon_{n}$. The sequence
$(s_{n},x_{n})_{n\in\mathbb{N}}$ corresponds to $(0,x_{n})_{n\in\mathbb{N}}$.
Interpreting $\phi_{n}$ and $\tilde{\phi}_{n}$ as skew products
$\pi_{n}$ and $\tilde{\pi}_{n}$ we see that $\tilde{\pi}_{n}\to\pi=\tau\times\pi_{0}$
which implies semi-singular convergence of $\pi_{n}$.
\end{example*}
Suppose now $S\subset X$ is an attractor for $\pi_{0}$. Then there
exists a stable isolating block $B$. Thus $N:=\mathbb{R}\times B$
is a stable isolating block for $\pi$ with $A_{\pi}(N)=\mathbb{R}\times S$.
Furthermore, it can be shown that for all $T>0$ \begin{eqnarray*}
G_{\pi}^{T}(N) & = & \mathbb{R}\times G_{\pi_{0}}^{T}(B)\\
& = & \mathbb{R}\times(B\pi_{0}T)\end{eqnarray*}
and $\Gamma_{\pi}^{T}(N)=\varnothing$.
Since $B$ is compact $(\cdot)\pi t:N\to N$ is uniformly continuous
for all $t\ge0$. And for each $T>0$ there is some $\delta>0$ such
that $U_{\delta}(G_{\pi}^{T}(N))\subset N$ where $U_{\epsilon}(A)=\{x\in X\,|\, d(x,A)<\epsilon\}$.
We assume here that $\pi_{n}$ does not blow-up in finite time in
$N$. This will always be the case if $\pi_{n}$ is a skew product
flow because the unbounded component of $N=\mathbb{R}\times B$ represents
the time and $B$ is compact. Hence whenever $\omega_{(s,x)}^{\pi_{n}}<\infty$
for $(s,x)\in N$ then $(s,x)\pi_{n}[0,\omega_{(s,x)}^{\pi_{n}})\not\subset N$.
\begin{lem}
For every $T>0$ there is an $n_{0}>0$ such that for all $n\ge n_{0}$\[
G_{\pi_{n}}^{T}(N)\ne\varnothing.\]
\end{lem}
\begin{proof}
easy exercise\end{proof}
\begin{lem}
\label{lem:Gamma_empty}There is a $T_{0}$ and a $n_{0}$ such that
for all $T\ge T_{0}$ and $n\ge n_{0}$\[
\Gamma_{\pi_{n}}^{T}(N)=\varnothing.\]
\end{lem}
\begin{rem*}
The proof is based on {\cite[Lemma 3.5]{Carbinatto2002}}. The idea
is to show that if it does not hold then $K=A_{\pi_{0}}(B)$ and $\partial B$
intersect non-trivial which is impossible because $K$ is isolated.
We will give the whole proof because the fact that $\tilde{z}\in A_{\pi_{0}}^{-}(B)$
will also apply to the infinite-dimensional case if we assume $\{\pi_{n}\}$-semi-singular-admissibility
of $N$. \end{rem*}
\begin{proof}
Suppose this is not the case. Then there exist a sequence $T_{n}\to\infty$
and a subsequence of $\pi_{n}$, also denoted by $\pi_{n}$, such
that $E_{n}:=\Gamma_{\pi_{n}}^{T_{n}}(N)\ne\varnothing$ for all $n$.
This implies that there is a sequence $x_{n}\in N$ such that for
some $s_{n}\in[0,T_{n}]$ we have $x_{n}\pi_{n}[0,t_{n}]\subset N$,
$x_{n}\pi_{n}T_{n}\in E_{n}$ and $x_{n}\pi t_{n}\in\partial N$ with
$t_{n}=T_{n}+s_{n}$.
Since $N=\mathbb{R}\times B$ we have $\partial N=\mathbb{R}\times\partial B$.
So that $x_{n}\pi t_{n}=(r_{n},\tilde{z}_{n})$ for some $\tilde{z}_{n}\in\partial B$.
Because $\partial B$ is compact there is a subsequence of $x_{n}\pi_{n}t_{n}$
denoted by $z_{n}^{0}=x_{n}^{0}\pi_{n}^{0}t_{n}^{0}$ such that $\tilde{z}_{n}^{0}\to\tilde{z}^{0}\in\partial B$.
Since $B$ is compact, $t_{n}^{0}\to\infty$ and $x_{n}^{0}\pi_{n}^{0}[0,t_{n}^{0}]\subset N$
there is a subsequence $(x_{n}^{1}\pi_{n}^{1}t_{n}^{1})_{n\in\mathbb{N}}$
with $t_{n}^{1}\ge1$ such that $(r_{n}^{1},\tilde{z}_{n}^{1})=x_{n}^{1}\pi_{n}^{1}(t_{n}^{1}-1)$
is defined and $\tilde{z}_{n}^{1}\to\tilde{z}^{1}\in B$. Recursively,
for each $k\ge1$ we get a subsequence $(x_{n}^{k}\pi_{n}^{k}t_{n}^{k})_{n\in\mathbb{N}}$
of $(x_{n}^{k-1}\pi_{n}^{k-1}t_{n}^{k-1})_{n\in\mathbb{N}}$ with
$t_{n}^{k}\ge k$ such that $z_{n}^{k}=(r_{n}^{k},\tilde{z}_{n}^{k})=x_{n}^{k}\pi_{n}^{k}(t_{n}^{k}-k)$
and $\tilde{z}_{n}^{k}\to\tilde{z}^{k}$ for some $\tilde{z}^{k}\in B$.
We claim that $\tilde{z}^{k}\pi_{0}t\in N$ all $t\in[0,k]$ and $\tilde{z}^{k}\pi_{0}k=\tilde{z}^{0}$,
i.e. $\tilde{z}^{0}\in A_{\pi_{0}}^{-}(B)$. Postponing the proof
of this claim we immediately get that $\tilde{z}^{0}\in\partial B\cap A^{-}(B)$
but $B$ is a stable isolating neighborhood such that $A(B)=A^{-}(B)\subset\operatorname{int}B$,
i.e. $\partial B\cap A_{\pi_{0}}^{-}(B)=\varnothing$, which is a
contradiction and thus the lemma is true.
It remains to proof our claim. Suppose $\tilde{z}^{k}\pi_{0}t_{0}\notin B$
for some $t_{0}\in[0,k]$. Because $\pi_{n}\overset{\mbox{\tiny ssing}}{\longrightarrow}\pi_{0}$
the sequence $\tilde{y}_{n}$ with $(s_{n},\tilde{y}_{n})=z_{n}^{k}\pi_{n}^{k}t_{0}$
converges to $\tilde{z}^{k}\pi_{0}t_{0}$, but $\tilde{y}_{n}\in B$
and $B$ is closed which is impossible. Hence $\tilde{z}^{k}\pi_{0}[0,k]\subset B$.
\end{proof}
Combining the two previous lemmas we obtain that there is a $T_{0}$
and an $n_{0}$ such that for all $T\ge T_{0}$ and $n\ge n_{0}$
the set $G_{\pi_{n}}^{T}(N)\ne\varnothing$ is a stable isolating
neighborhood for $\pi_{n}$.
Because $B$ is compact and $(\cdot)\pi t$ is uniformly continuous
on $N$, there is a $\delta=\delta(\pi_{0},B)>0$ such that for $T\ge T_{0}$
\begin{eqnarray*}
U_{\delta}(G_{\pi}^{2T}(N)) & \subset & G_{\pi}^{T}(N)\end{eqnarray*}
and \begin{eqnarray*}
U_{\delta}(G_{\pi}^{T}(N)) & \subset & N.\end{eqnarray*}
Furthermore, choose $\delta$ small enough and $T$ large enough so
that $U_{\delta}(B)\subset B'$ (see remark after theorem \ref{thm:stable})
and \[
G_{\pi_{0}}^{T}(B')\subset B,\]
i.e. $U_{\delta}(G_{\pi}^{T}(U_{\delta}(N)))\subset N$ holds as well.
\begin{rem*}
Benci (\cite{Benci1991}) defines a set $\Sigma_{0}$ which contains
all sets having a similar property and uses it to prove the continuation
of the Conley index. Most proofs of the inclusions like $U_{\delta}(G_{\pi}^{2T}(N))\subset G_{\pi_{n}}^{T}(N)$
are omitted in that paper.
\end{rem*}
In addition assume now that each $\pi_{n}$ is a skew product flow,
i.e. \[
P_{1}((s,x)\pi_{n}t)=s+t.\]
\begin{thm}
\label{thm:finite}If for large $n\ge n_{0}$ and $T\ge T_{0}$ we
have \[
d(x\pi_{n}t,x\pi t)<\frac{\delta}{3}\]
for all $t\in[-T,T]$ and $x\in N$ then\[
A_{\pi_{n}}(N)\ne\varnothing\]
is an isolated invariant set with compact $t$-slices such that for
an $\epsilon=\epsilon(\pi_{0},B,T)>0$ \[
U_{\epsilon}(A_{\pi_{n}}(N))\subset G_{\pi_{n}}^{T}(N).\]
\end{thm}
\begin{rem*}
The theorem shows that $A_{\pi_{n}}(N)$ is a past attractor in the
sense of Rasmussen \cite{Rasmussen2007}.\end{rem*}
\begin{proof}
Let $x\in G_{\pi_{n}}^{T}(N)$ then \[
x\pi_{n}[-T,T]\subset N.\]
Since $d(x\pi t,x\pi_{n}t)<\delta$ for all $t\in[-T,T]$ we have
\[
x\pi[-T,T]\subset U_{\delta}(N).\]
This implies that $G_{\pi_{n}}^{T}(N)\subset G_{\pi}^{T}(U_{\delta}(N))$
and thus $U_{\delta}(G_{\pi_{n}}^{T}(N))\subset N$ because $U_{\delta}(G_{\pi}^{T}(U_{\delta}(N)))\subset N$
Because $\pi$ is uniformly continuous on $N\times[-T,T]$ we can
choose $\epsilon>0$ such that $d(x,\tilde{x})<\epsilon$ implies
$d(x\pi t,\tilde{x}\pi t)<\frac{\delta}{3}$ for all $x,\tilde{x}\in N$
and $t\in[-T,T]$. Now if $x\in U_{\epsilon}(G_{\pi_{n}}^{2T}(N))$
then there is an $\tilde{x}\in G_{\pi_{n}}^{2T}(N)$ such that $d(x,\tilde{x})<\epsilon$
and \[
\tilde{x}\pi_{n}[-T,T]\subset G_{\pi_{n}}^{T}(N).\]
Furthermore, we have for all $t\in[-T,T]$ \begin{eqnarray*}
d(x\pi_{n}t,\tilde{x}\pi_{n}t) & \le & d(x\pi_{n}t,x\pi t)+d(x\pi t,\tilde{x}\pi t)+d(\tilde{x}\pi t,\tilde{x}\pi_{n}t)\\
& < & \frac{\delta}{3}+\frac{\delta}{3}+\frac{\delta}{3}=\delta.\end{eqnarray*}
Since $U_{\delta}(G_{\pi_{n}}^{T}(N))\subset N$ we have $x\pi_{n}[-T,T]\subset N$,
i.e. $U_{\epsilon}(G_{\pi_{n}}^{2T}(N))\subset G_{\pi_{n}}^{T}(N)$.
In particular, $U_{\epsilon}(A_{\pi_{n}}(N))\subset G_{\pi_{n}}^{T}(N)$.
Now let $x\in G_{\pi}^{2T}(N)$ then $x\pi[-T,T]\subset G_{\pi}^{T}(N)$.
Because $U_{\delta}(G_{\pi_{n}}^{T}(N))\subset N$ and $d(x\pi t,x\pi_{n}t)<\delta$
for all $t\in[-T,T]$ we have $x\in G_{\pi_{n}}^{T}(N)$. Hence \[
G_{\pi}^{2T}(N)=\mathbb{R}\times G^{2T}(B)\subset G_{\pi_{n}}^{T}(N).\]
Since $G_{n}:=G_{\pi_{n}}^{T}(N)$ is closed the set $G_{n}(t)=G_{\pi_{n}}^{T}(N)\cap\{t\}\times X\subset\{t\}\times B$
is compact and non-empty for each $t\in\mathbb{R}$. Now define\[
A(t)=\bigcap_{t\ge0}G_{n}(-t)\pi_{n}t\subset\{t\}\times B.\]
$G_{n}$ is positive invariant so that this intersection is a decreasing
sequence of non-empty compact sets and thus compact and non-empty
itself. Let \[
A=\bigcup_{t\in\mathbb{R}}A(t).\]
Then $A$ is invariant w.r.t. $N$ and $A(N)\subset A\subset N$ which
implies $A(N)=A\ne\varnothing$.
\end{proof}
\subsection{\label{sub:non-auto}Non-autonomous dynamical systems}
For a general NDS $(\phi,\theta)$ we do not have $P=\mathbb{R}$.
Assume $X$ is locally compact and $(\phi,\theta)$ is a two-sided
process. Denote by $P^{*}$ the set of $\theta$-orbits of $P$. Then
each orbit $\sigma\in P^{*}$ is represented by a $p\in P$. So after
choosing one representative for each orbit we can assume w.l.o.g.
$P^{*}\subset P$.
Because $(\phi,\theta)$ is two-sided for each $p\in P^{*}$ we have
an invertible skew product flow $\pi_{p}$ with \[
(s,x)\pi_{p}t=(s+t,x\phi^{p\theta s}t).\]
Suppose $B$ is a stable isolating neighborhood of $\pi_{0}$. If
each $\pi_{p}$ is close to $\pi=\tau\times\pi_{0}$, i.e. there is
a sufficiently small $\delta>0$ and a $T>0$ (both independent of
$p$) such that \[
d((s,x)\pi t,(s,x)\pi_{p}t)<\delta\]
for all $t\in[-T,T]$ and $x\in B$ then theorem \ref{thm:finite}
implies \[
G_{p}=G^{T}(\mathbb{R}\times B)\]
is a stable isolating neighborhood for $\pi_{p}$ and their isolated
invariant sets $A_{p}=A_{\pi_{p}}(\mathbb{R}\times B)$ are non-empty.
Now define the set-valued map $D:P\to2^{X}$ by \[
D(q)=\{x\in X\,|\,\mbox{for some \ensuremath{s\in\mathbb{R}}, \ensuremath{p\in P^{*}}\,\ s.t. \ensuremath{(s,x)\in G_{p}}\,\ and }p\theta s=q\}\]
It is an easy exercise to show that $D$ maps into the set of closed
sets and is forward invariant w.r.t. the non-autonomous system $(\phi,\theta)$.
In the same way we can show that there is set-valued map $A$ which
is invariant and \[
A(q)=P_{2}((A_{p}\pi_{p}s)\cap\{s\}\times X)\]
is compact for $q=p\theta s$ and $p\in P^{*}$. Furthermore, the
$\epsilon$ in theorem \ref{thm:finite} only depends on $T$, $B$
and $\pi_{0}$, i.e. \[
U_{\epsilon}(A(q))\subset D(q)\]
which shows that $A$ is a past attractor.
\subsection{\label{sub:rds}Random dynamical systems}
A random dynamical system is an NDS such that $P=\Omega$ is a probability
space, $\theta$ is measurable dynamical system and the map\[
(t,\omega)\mapsto x\phi^{\omega}t\]
is measurable for every $x$ in $X$.
If we show that the sets $D$ and $A$ are random closed sets then
$A$ is a random past attractor. And we get:
\begin{thm*}
Sufficiently small (bounded) noise does not destroy local attractors.
\end{thm*}
In particular, if the global attractor of the unperturbed system has
several disjoint local attractors ({}``sinks'') then the same holds
for the global attractor of the perturbed system. This is a complementary
result to Crauel, Flandoli - {}``Additive Noise Destroys a Pitchfork
Bifurcation'' \cite{Crauel1998} where they show that the {}``perturbed''
global attractor is a single point with probability $1$ when {}``small''
white noise is applied.
It remains to show that $D$ and $A$ are closed random sets. It suffices
to show that $D$ is random compact. Each $G_{\omega}^{T}(N)$ is
flow-defined, i.e.\[
G_{\omega}^{T}(N)=\bigcap_{t\in[-T,T]}(\mathbb{R}\times B)\pi_{\omega}t.\]
Because the $\pi_{\omega}$ are skew products, for each $s$-slice
we have \begin{eqnarray*}
G_{\omega}^{T}(N)(s) & = & G_{\omega}^{T}(N)\cap(\{s\}\times X)\\
& = & \{s\}\times\bigcap_{t\in[-T,T]}(B\phi^{\omega\theta(s-t)}t).\end{eqnarray*}
So that for $\varpi=\omega\theta s$ \begin{eqnarray*}
D(\varpi) & = & \bigcap_{t\in[-T,T]}(B\phi^{\varpi\theta(-t)}t)\\
& = & \bigcap_{t\in[-T,T]\cap\mathbb{Q}}(B\phi^{\varpi\theta(-t)}t)\end{eqnarray*}
Because $B$ is a deterministic compact set, $\phi$ invertible and
$D$ a decreasing intersection of compact sets, $D$ is a random compact
set. In particular, we see that the choice of the representative $\omega\in\Omega^{*}=P^{*}$
does not matter and each $D(\varpi)$ is defined in the same way.
\section{\label{sec:Inf}Infinite dimensional case}
In this section we show that the index pair continuation in {\cite[Theorem 12.3]{Rybakowski1987}
still holds if we replace the admissibility arguments by semi-admissibility
and semi-singular convergence arguments. We are not able to show that
two index pairs of the perturbed (non-autonomous) system are Conley
equivalent, i.e. the quotients have the same homotopy type in the
category of pointed space. So the Conley index is not well-defined.
Even worse, because the index pair is unbounded the Wa{\.z}ewski
principle does not imply that the invariant set is non-empty if the
quotient space is non-trivial. Nevertheless, we can use the continuation
to show that a stable isolating neighborhood continues to a stable
isolating neighborhood, i.e. the exit set of the perturbed system
is empty.
\begin{rem*}
As mentioned in the introduction appropriate versions of theorem \ref{thm:cont}
and corollary \ref{cor:cont} hold also for stable isolating neighborhoods
in the discrete time setting using the ideas of \cite{Kell2011c}.
The definitions of semi-(singular-)admissibility and semi-singular
convergence can be given similarly.
Furthermore, for both, continuous and discrete time, instead of a
semi-admis\-sibility of the perturbed maps we could use {}``weak
properness'' of the space $X$, i.e. bounded $\delta$-neighborhoods
of compact sets are weakly compact, to show that the invariant set
is non-empty and a weak attractor (see \cite{Kell2011c} for definition).\end{rem*}
\begin{defn}
[Semi-singular admissiblity] Let $\pi_{n}$ be semiflows on $\mathbb{R}\times X$
and $\pi_{0}$ be a semiflow on $X$ such that $\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$.
A closed subset $N$ of $\mathbb{R}\times X$ is called $\{\pi_{n}\}$-semi-singular-admissible
(ss-admissible) if the following holds:
\begin{itemize}
\item Let $(s_{n})_{n\in\mathbb{N}}$ be a sequence in $\mathbb{R}$. If
$x_{n}\in X$ and $t_{n}\in\mathbb{R}^{+}$ with $0\le t_{n}<\omega_{(s_{n},x_{n})}^{\pi_{n}}$
are two sequences such that $t_{n}\to\infty$ as $n\to\infty$ and
$(s_{n},x_{n})\pi_{n}[0,t_{n}]\subset N$ then the sequence $\{P_{2}((s_{n},x_{n})\pi_{n}t_{n})\}_{n\in\mathbb{N}}$
of the projected endpoints has a convergent subsequence.
\end{itemize}
If, in addition, each $\pi_{n}$ does not explode in $N$, i.e. \[
(s,x)\pi_{n}[0,\omega_{(s,x)}^{\pi_{n}})\subset N\:\mbox{ implies }\:\omega_{(s,x)}^{\pi_{n}}=\infty\]
then we say that $N$ is strongly $\{\pi_{n}\}$-ss-admissible.
\end{defn}
The definition means that $\pi_{n}$ should not converge too badly
to $\pi_{0}$. It is a property of the sequence $\{\pi_{n}\}_{n\in\mathbb{N}}$
and does not tell anything about $\pi_{0}$ alone and neither about
a single $\pi_{n}$. An adjusted admissibility definition of a single
semiflow $\pi$ for our setting is given as follows.
\begin{defn}
[Semi-Admissiblity] Let $\pi$ be a semiflow on $\mathbb{R}\times X$.
A closed subset $N$ of $\mathbb{R}\times X$ is called $\pi$-semi-admissible
if the following holds:
\begin{itemize}
\item Let $(s_{n})_{n\in\mathbb{N}}$ be a sequence in $\mathbb{R}$. If
$x_{n}\in X$ and $t_{n}\in\mathbb{R}^{+}$ with $0\le t_{n}<\omega_{(s_{n},x_{n})}$
are two sequences such that $t_{n}\to\infty$ as $n\to\infty$ and
$\{s_{n}+t_{n}\}_{n\in\mathbb{N}}$ is precompact, $(s_{n},x_{n})\pi[0,t_{n}]\subset N$
then the sequence of endpoints $\{(s_{n},x_{n})\pi t_{n}\}_{n\in\mathbb{N}}$
has a convergent subsequence.
\end{itemize}
If, in addition, $\pi$ does not explode in $N$, i.e. \[
(s,x)\pi[0,\omega_{(s,x)})\subset N\:\mbox{ implies }\:\omega_{(s,x)}=\infty\]
then we say that $N$ is strongly $\pi$-semi-admissible.\end{defn}
\begin{rem*}
For a semiflow $\pi_{0}$ on $X$ the usual strong $\pi_{0}$-admissibility
of a closed subset $N$ is defined without the sequence $s_{n}$ (see
{\cite[4.1]{Rybakowski1987}}).
\end{rem*}
Suppose $B$ is closed and strongly $\pi_{0}$-admissible, $B\subset U$
for some open $U\subset X$ with strong $\pi_{0}$-admissible closure
and $d(B,\partial U)\ge\delta$. We say that $\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$
converges uniformly if $d(P_{2}((s,x))\pi_{n}t^{*},x\pi_{0}t^{*})<\epsilon_{n,t}\to0$
for $t>0$ and all $(s,x)\in\mathbb{R}\times U$ with $t<\min\{\omega_{x}^{\pi_{0}},\omega_{(s,x)}^{\pi_{n}}\}$.
\begin{rem*}
Uniform convergence of $\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$
is similar to the assumption of Benci \cite{Benci1991} used to prove
his continuation theorem for the Conley index. \end{rem*}
\begin{lem}
Suppose $\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$
uniformly and $\pi_{n}$ does not explode in $N=\mathbb{R}\times B$.
Then $N$ is strongly $\{\pi_{n}\}$-admissible.\end{lem}
\begin{proof}
Let $(s_{n},x_{n})\in N$ and $t_{n}\to\infty$ be a sequence fulfilling
the assumption of the definition of ss-admissibility. Because of uniform
convergence and $d(N,\partial U)\ge\delta$, there is an $N(t)$ for
each such that for all $n\ge N(t)$\[
x\pi_{0}[0,t]\subset U\]
whenever $(s,x)\in N$ with $(s,x)\pi_{n}t\subset N$.
Thus there is a sequence $r_{n}\ge0$ with $t_{n}-r_{n}\to\infty$
such that $(s_{n},x_{n})\pi_{n}[0,t_{n}]\subset N$ implies that \[
y_{n}\pi_{0}[0,t_{n}-r_{n}]\subset U\]
for $(\tilde{s}_{n},y_{n})=(s_{n},x_{n})\pi_{n}r_{n}$. Furthermore,
we can choose $r_{n}$ such that $\delta_{n}=\max_{t\in[0,t_{n}-r_{n}]}\epsilon_{n,t}\to0$
and therefore \[
d(P_{2}((s_{n},x_{n})\pi_{n}t_{n})),y_{n}\pi(t_{n}-r_{n}))\le\delta_{n}.\]
Because the closure of $U$ is strongly $\pi_{0}$-admissible, the
sequence of endpoints $\{y_{n}\pi(t_{n}-r_{n})\}_{n\in\mathbb{N}}$
has a convergent subsequence which implies that $\{P_{2}((s_{n},x_{n})\pi_{n})\}_{n\in\mathbb{N}}$
has a convergent subsequence, in particular the limit point is in
$N$.\end{proof}
\begin{defn}
[index pair]Let $\tilde{N}$ be a closed isolating neighborhood for
$K=A(\tilde{N})$. A pair $(N,L)$ is called an index pair in $\tilde{N}$
(w.r.t. $\pi$) if $L\subset N\subset\tilde{N}$ and the following
holds
\begin{enumerate}
\item $K\subset\operatorname{int}(N\backslash L)$ and $K$ is the maximal
invariant set in $\operatorname{cl}(N\backslash L)$
\item if $x\in L$ and $x\pi[0,\epsilon]\subset N$ then $x\pi[0,\epsilon]\subset L$,
i.e. $L$ is positive $\pi$-invariant relative to $N$
\item if $x\in N$ and $x\pi[0,\omega_{x})\not\subset N$ then there is
a $0\le t<\omega_{x}$ such that $x\pi t\in L$, i.e. $L$ is an exit
ramp for $N$
\end{enumerate}
\end{defn}
The following theorem is a variant of {\cite[I-12.3-12.7]{Rybakowski1987}}
(see also \cite{Carbinatto2002}) by replacing all admissibility arguments
with semi-singular-admissibility argument and replacing $g^{-}:\tilde{N}\to\mathbb{R}^{\ge0}$
by a lifted version $\tilde{g}^{-}:\mathbb{R}\times\tilde{N}\to\mathbb{R}^{\ge0}$
defined by $\tilde{g}^{-}(s,x)=g^{-}(x)$.
\begin{thm}
\label{thm:cont}Let $\pi_{0}$ be a local semiflow on $X$ and $\pi_{n}$,
$n\in\mathbb{N}$, be local semiflows on $\mathbb{R}\times X$. Suppose
$\tilde{N}$ is a closed set in $X$ and strongly $\pi_{0}$-admissible.
Moreover, assume $\pi_{n}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$
as $n\to\infty$ and $\mathbb{R}\times\tilde{N}$ is strongly $\{\pi_{n_{m}}\}$-semi-admissible
for all subsequences $\{\pi_{n_{m}}\}_{m\in\mathbb{N}}$. Set $K_{n}=A_{\pi_{n}}(\mathbb{R}\times\tilde{N})$,
$K_{0}=A_{\pi}(\mathbb{R}\times\tilde{N})$ and assume $K=A_{\pi_{0}}(\tilde{N})\subset\operatorname{int}\tilde{N}$.
Then there exist an $n_{0}\in\mathbb{N}$ and two closed set $N$
and $N^{'}$ such that for each $n\in\{n\ge n_{0}\}\cup\{0\}$ there
are two index pair $\langle N_{n},L_{n}\rangle$ and $\langle N_{n}^{'},L_{n}^{'}\rangle$
with the following properties:
\begin{itemize}
\item $K_{n}\subset N^{'}\subset\operatorname{int}(\mathbb{R}\times N)\subset\operatorname{int}(\mathbb{R}\times\tilde{N})$,
$N_{0}^{'}=\mathbb{R}\times N^{'}$ and $N_{0}=\mathbb{R}\times N$
\item $\langle N_{n}^{'},L_{n}^{'}\rangle$ (resp. \textup{$\langle N_{n},L_{n}\rangle$)
is an index pair in $N_{0}^{'}$ (resp. in $N_{0}$) w.r.t $\pi_{n}$
(resp. w.r.t $\pi=\tau\times\pi_{0}$ if $n=0$)}
\item $N_{n}^{'}\subset N_{0}^{'}\subset N_{n}\subset N_{0}$ and $L_{n}^{'}\subset L_{0}^{'}\subset L_{n}\subset L_{0}$.
\end{itemize}
\end{thm}
\begin{proof}
Since the proof will follow the original one almost completely and
we are only interested in attractors, we only show a proof in case
$\tilde{N}$ is positive $\pi_{0}$-invariant. In particular, we will
assume that $\tilde{N}$ is an isolating block with empty exit set
defined via the function $g^{-}$. For $a>0$ define \[
V(a)=\{x\in\tilde{U}\,|\, g^{-}(x)<a\}.\]
Then there is an $a_{0}>0$ such that $N:=\operatorname{cl}V(a_{0})\subset\tilde{U}$.
And similar to {\cite[I-4.5]{Rybakowski1987}} we can show that for
$0<\epsilon\le a_{0}$ and all $n\ge n_{0}(\epsilon)$ \[
K_{n}\subset\mathbb{R}\times V(\epsilon).\]
Define \[
N_{n}(\epsilon)=(\mathbb{R}\times N)\cap\operatorname{cl}\{\tilde{y}\,|\,\mbox{\ensuremath{\exists\tilde{x}\in\mathbb{R}\times V(\epsilon)}, \ensuremath{t\ge0}\,\ s.t.\,\ \ensuremath{\tilde{x}\pi_{n}[0,t]\subset\mathbb{R}\times\tilde{U}\,}and \ensuremath{\tilde{x}\pi_{n}t=\tilde{y}}}\}.\]
Following the proof of {\cite[I-12.5]{Rybakowski1987}} we can show
that $N_{n}(\epsilon)$ satisfies the following properties for $n\ge n_{0}(\epsilon)$
\begin{itemize}
\item $x\in N_{n}(\epsilon)$ and $x\pi_{n}[0,t]\subset\mathbb{R}\times N$
implies $x\pi_{n}t\in N_{n}(\epsilon)$
\item $K_{n}\subset\mathbb{R}\times V(\epsilon)\subset N_{n}(\epsilon)$
\end{itemize}
We claim that for small $\epsilon_{0}>0$ whenever $\epsilon\le\epsilon_{0}$
and $n\ge n_{0}(\epsilon)$ then $N_{n}(\epsilon)$ is positive $\pi_{n}$-invariant.
If this is not true then there is a sequence $\epsilon_{m}\to0$ and
\[
y_{m}=(s_{m},z_{m})\in N_{n_{m}}(\epsilon_{m})\cap(\mathbb{R}\times\partial N).\]
By definition of $N_{n_{m}}(\epsilon_{m})$ there is a sequence $\tilde{y}_{m}\in X$,
$x_{m}\in\mathbb{R}\times V(\epsilon_{m})$ and $t_{m}\ge0$ such
that $d(y_{m},\tilde{y}_{m})<2^{-m}$, $x_{m}\pi_{n_{m}}[0,t_{m}]\subset\tilde{U}$
and $\tilde{y}_{m}=x_{m}\pi_{n_{m}}t_{m}$. Because $g^{-}(x_{m})\to0$
and $A_{f}^{-}(B)=A_{f}(B)$ we can assume w.l.o.g. that $x_{m}\to x_{0}\in A_{f}(B)$.
Admissibility and $\pi_{n_{m}}\overset{\mbox{\mbox{\tiny ssing}}}{\longrightarrow}\pi_{0}$
imply the sequence $\{P_{2}(x_{m}\pi_{n_{m}}t_{m})\}_{m\in\mathbb{N}}$
has a convergent subsequence and w.l.o.g. $P_{2}(\tilde{y}_{m})=P_{2}(x_{m}\pi_{n_{m}}t_{m})\to z_{0}\in A_{f}^{-}(\tilde{N})=A_{f}(\tilde{N})\subset\operatorname{int}N$
(see proof of lemma \ref{lem:Gamma_empty}) and thus $P_{2}(y_{m})\to z_{0}$.
Since $P_{2}(z_{m})\in\partial N$ we must have $y_{0}\in\partial N$,
but this is a contradiction since $\partial N\cap A_{\pi_{0}}(\tilde{N})=\varnothing$.
Set $N^{'}=\operatorname{cl}V(\epsilon)$ which is also positive $\pi_{0}$-invariant.
Applying the arguments above for $N'$ we get another positive $\pi_{n}$-invariant
$N_{n}^{'}\subset\mathbb{R}\times N^{'}$. So we have \[
N_{n}^{'}\subset\mathbb{R}\times N^{'}\subset N_{n}\subset\mathbb{R}\times N.\]
Positive invariance and $K_{n}\subset\mathbb{R}\times V(\delta)\subset N_{n}^{'}\subset N_{n}$
for some $\delta>0$ implies that $N_{n}$ and $N_{n}^{'}$ are index
pairs.\end{proof}
\begin{cor}
\label{cor:cont}If, in addition to the assumption of the theorem,
we assume that each $\pi_{n}$ is skew product flows and $\tilde{N}$
is a stable isolating neighborhood for $\pi_{0}$ and $N_{0}=\mathbb{R}\times N$
from the theorem is strongly $\pi_{n}$-semi-admissible then $K_{n}\ne\varnothing$
and each $t$-slice $K_{n}\cap\{t\}\times X$ is non-empty and compact.
Furthermore, there is an $n_{0}=n_{0}(\tilde{N},\pi_{0})$ and a $\delta=\delta(\tilde{N},\pi_{0})>0$
such that for all $n\ge n_{0}$ \[
U_{\delta}(K_{n})\subset N_{n},\]
i.e. $K_{n}$ is a past attractor in the sense of Rasmussen \cite{Rasmussen2007}.
And for the sequence $(K_{n})_{n\in\mathbb{N}}$ we have \[
\lim_{n\to\infty}\sup_{y\in K_{n}}\inf_{x\in K}d(P_{2}y,x)=0,\]
i.e. $K_{n}$ is upper-semicontinuous {}``at $K_{0}$'' as $n\to\infty$.\end{cor}
\begin{rem*}
In \cite{Caraballo2003} an upper-semicontinuity is proved by assuming
that a global attractor exists. Our result shows that the same is
true for local attractors. We can prove without further assumptions
on the system that such local attractors always exist if the perturbation
is {}``uniformly small''. If $K$ is the global attractor of $\pi_{0}$
and the assumption (h2) of \cite{Caraballo2003} holds then $K_{n}$
must be the global attractor for large $n$.\end{rem*}
\begin{proof}
If $\tilde{N}$ is stable then w.l.o.g. we can replace $\tilde{N}$
by a stable isolating block defined via the function $g^{-}$ (see
below). The previous theorem implies that $N_{n}$ is a stable isolating
neighborhood for $\pi_{n}$ and $n\ge n_{0}$, i.e. $N_{n}=A_{\pi_{n}}^{+}(N_{n})$.
Because $N_{0}^{'}=\mathbb{R}\times N'\subset N_{n}$ we have \[
N_{n}\cap\mathbb{R}\times X\ne\varnothing.\]
Thus there is a sequence $s_{k}\to-\infty$ and $(x_{k})_{k\in\mathbb{N}}$
such that $(s_{k},x_{k})\in N_{n}$. Because $N_{n}$ is positive
$\pi_{n}$-invariant and $N_{0}$ is strongly $\pi_{n}$-semi-admissible
the sequence \[
(s_{k},x_{k})\pi_{n}(-s_{k})\in N_{n}\cap\{0\}\times X\]
has a convergent subsequence and each limit point is in $A_{n}^{-}(N_{n})$
and thus in $A_{\pi_{n}}(N_{n})\subset\operatorname{int}N_{n}^{'}$.
By the same argument we can show that each $t$-slice of $N_{n}$
contains a non-empty compact $t$-slice of the invariant set.
The upper-semicontinuity is a standard result from the index continuation
(see e.g. {\cite[Corollary 4.11]{Carbinatto2002}}).
It remains to show that there is a $\delta>0$ such that for large
$n$\[
U_{\delta}(K_{n})\subset N_{n}.\]
Recall the definition of the function $g^{-}:\tilde{N}\to\mathbb{R}^{\ge0}$
(adjusted to the stable case)\[
g^{-}(x):=\sup\{\alpha(t)F(x\pi t)\,|\,0\le t<\infty\}\]
with \[
F(x):=\min\{1,\operatorname{dist}(x,K)\}\]
and some strictly increasing $C^{\infty}$-diffeomorphism $\alpha:[0,\infty)\to[1,2)$.
$g^{-}$ is continuous on $\tilde{N}$ and strictly decreasing along
$\pi_{0}$ outside of $K$. For sufficiently small $0<\epsilon<\frac{1}{2}$
\[
B_{\epsilon}:=(g^{-})^{-1}([0,\epsilon])\subset\operatorname{int}\tilde{N}\]
defines a stable isolating block and $B_{\delta}\subset B_{\epsilon}$
for $\delta<\epsilon$. Since $g^{-}(x)<1$ for $x\in B$ we have
$\operatorname{dist}(x,K)\le g^{-}(x)$. Because $K$ is compact there
exists some $2\delta<\epsilon$ such that for all $x\in\partial B$
\[
2\delta<d(x,K)\le\epsilon.\]
This implies \[
U_{\delta}(B_{\delta})\subset B_{\epsilon}.\]
The same applies for the suspension $\tilde{g}^{-}:\mathbb{R}\times\tilde{N}\to\mathbb{R}^{\ge0}$.
Our proof of the previous theorem constructs the sets $N_{0}^{'}$
and $N_{0}$ from the function $g^{-}$ (resp. $\tilde{g}^{-}$ in
our adapted version), i.e. \begin{eqnarray*}
N_{0}^{'} & = & (\tilde{g}^{-})^{-1}([0,\epsilon])\\
& = & \mathbb{R}\times(g^{-})^{-1}([0,\epsilon])\\
& = & \mathbb{R}\times B_{\epsilon}\end{eqnarray*}
for some $\epsilon>0$. By the previous argument there is a $\delta>0$
such that for $N_{0}^{''}:=(\tilde{g}^{-})^{-1}([0,\delta])$ we have
\[
U_{\delta}(N_{0}^{''})\subset N_{0}^{'}.\]
Similar arguments as used in the proof above show that there is a
stable isolating neighborhood $N_{n}^{''}$ of $K_{n}=A_{\pi_{n}}(\mathbb{R}\times\tilde{N})$
with \[
N_{n}^{''}\subset N_{0}^{''}\subset N_{n}^{'}\subset N_{0}^{'}\subset N_{n}\subset N_{0}.\]
Now the inclusion sequence implies \[
U_{\delta}(K_{n})\subset U_{\delta}(N_{n}^{''})\subset N_{n}.\]
\end{proof}
\subsection{Retarded functional differential equations}
Let $C=C([-r,0],\mathbb{R}^{m})$ be the space of continuous functions
equipped with the $\sup$-norm, $r\ge0$ and $\Omega\subset C$ be
an open set. For a continuous map $x:[-r+t,t]\to\mathbb{R}^{m}$ and
$t\in\mathbb{R}$ we write $x_{t}$ as the element in $C$ such that
$x_{t}(\theta)=x(t+\theta)$ for $\theta\in[-r,0].$
If $f_{0}:\Omega\to\mathbb{R}^{m}$ is Lipschitz continuous then the
following (autonomous) retarded functional differential equation (RFDE)
\begin{eqnarray*}
\dot{x}(t) & = & f_{0}(x_{t})\end{eqnarray*}
induces a semiflow $\pi_{0}$ such that each bounded and closed $N\subset\Omega$
for which $f_{0}(N)$ is bounded is strongly $\pi_{0}$ admissible
(see \cite{Hale1993,Rybakowski1987}).
\begin{rem*}
An RFDE with $r=0$ is an ODE on $\Omega\subset\mathbb{R}^{m}$.
\end{rem*}
Similarly if $f_{n}:\mathbb{R}\times\Omega\to\mathbb{R}^{m}$ is Lipschitz
continuous then the non-autonomous RFDEs \[
\dot{x}(t)=f_{n}(t,x_{t})\]
induce (skew product) semiflows $\pi_{n}$ on $\mathbb{R}\times\Omega$.
Furthermore, $\pi_{n}$ does not blow up in $\mathbb{R}\times N$
if $N\subset\Omega$ is closed and bounded and $f_{n}(t,N)$ is bounded
uniformly in $t$. One can even show that $\mathbb{R}\times N$ is
strongly $\pi_{n}$-semi-admissible (see proof of {\cite[4.2]{Rybakowski1987}}).
If we assume that \[
\sup_{t\in\mathbb{R}}\|f_{n}(t,\cdot)-f_{0}\|<\epsilon_{n}\to0\]
then we can easily show that $\pi_{n}\overset{\tiny\mbox{ssing}}{\longrightarrow}\pi_{0}$.
$\mathbb{R}\times N$ is strongly $\{\pi_{n}\}$-semi-singular-admissible.
\subsection{Semilinear parabolic equations}
The idea for semilinear parabolic equations is very similar. Suppose
$A$ is a sectorial operator and $f_{0}:X^{\alpha}\to X$ and $\Omega$
are {}``nice'' then \begin{eqnarray*}
u_{t} & = & Au+f_{0}(u)\\
u|_{\partial\Omega} & = & 0\end{eqnarray*}
induces a local semiflow $\pi_{0}$ on some $X^{\alpha}$. Furthermore,
bounded closed set $N\subset X^{\alpha}$ with $f_{0}(N)$ bounded
are strongly $\pi_{0}$-admissible.
As above the non-autonomous equation \begin{eqnarray*}
u_{t} & = & Au+f_{n}(t,u)\\
u|_{\partial\Omega} & = & 0\end{eqnarray*}
induces a skew product semiflow on $\mathbb{R}\times X^{\alpha}$
and $\mathbb{R}\times N$ is strongly $\pi_{n}$-semi-admissible for
a bounded closed $N\subset X^{\alpha}$ with $f_{n}(t,N)$ is bounded
uniformly in $t$. If, furthermore, \[
\sup_{t\in\mathbb{R}}\|f_{n}(t,\cdot)-f\|<\epsilon_{n}\to0\]
then $\pi_{n}\overset{\tiny\mbox{ssing}}{\longrightarrow}\pi_{0}$
and $\mathbb{R}\times N$ is strongly $\{\pi_{n}\}$-semi-singular-admissible.
\bibliographystyle{amsalpha}
|
\section{}
\newcommand{\mathrm{d}}{\mathrm{d}}
\def\kern-.2em\r{}\kern-.3em{\kern-.2em\r{}\kern-.3em}
Precise knowledge of electronic structures near the Fermi level $E_F$ is a prerequisite for understanding the origins of novel superconductivity. In the case of the iron-pnictide superconductors\cite{Kamihara08JACS}, ordinary electronic band structure calculations of the parent compounds greatly overestimate the antiferromagnetic (AFM) ordered moment\cite{Yin08PRL, Mazin09NatPhys, Ishibashi08JPSJS}. Several moment-reduction methods have been proposed, e.g., pnictogen-height adjustment\cite{Mazin09NatPhys}, negative-$U$\cite{Nakamura08condmat}. Moreover, a dynamical mean-field theory (DMFT) study suggests that correlations seriously modify the Fermi surface (FS) in the AFM phase\cite{Yin11NatPhys}. There is a clear need for an experimentally determined benchmark FS for judging how successfully those theoretical approaches reproduce band structures.
Recently, it was shown that BaFe$_2$As$_2$ can be detwinned mechanically by compressing or elongating single crystals along a tetragonal [110] axis\cite{Chu10Science, Tanatar10PRB}. This allows one to resolve intrinsic in-plane anisotropy in the AFM phase. In this Letter, we completely determine the FS in the AFM phase of BaFe$_2$As$_2$ via Shubnikov-de Haas (SdH) oscillation measurements on detwinned single crystals. The determined FS is reasonably accounted for by our band structure calculation but appreciably differs from the FS observed by previous angle-resolved photoemission spectroscopy (ARPES) studies \cite{LiuC09PRL, Yang09PRL, Malaeb09JPSJ, Richard10PRL, Yi11PNAS}. We also consider implications of the present results for transport properties.
High-quality BaFe$_2$As$_2$ single crystals with the residual resistivity ratios of 40--60 were prepared by a self-flux method and subsequent anneal\cite{Nakajima11PNAS}. A device similar to one described in Ref.~\onlinecite{Nakajima11PNAS} was used to mechanically detwin them. Standard four-contact resistivity $\rho$ measurements were performed in a dilution refrigerator and superconducting magnet. For a sample compressed (elongated) along a tetragonal [110] axis, the electrical current was applied parallel to the compression (elongation) direction, resulting in $I \parallel b$ ($I \parallel a$) geometry [Fig. 1(b)]. The electronic band structure of BaFe$_2$As$_2$ was calculated within the local spin-density approximation (LSDA) using a full-potential LAPW method (TSPACE and KANSAI-06). The experimental crystal structure at $T$ = 20 K \cite{Rotter08PRB} was used. The AFM order of the Fe moments was incorporated using the space group $Cccm$. The calculated magnetic moment is 1.6 $\mu_B$/Fe.
\begin{figure*}
\includegraphics[width=16cm]{v3fig1.pdf}
\caption{\label{fig1}(color online). (a) $T$-dependence of $\rho$ near the structural/magnetic phase transition. (b) Schematic diagram of the compressed sample. (c) $B$-dependence of $\rho$ in the compressed sample for three principal field directions at $T \sim$0.17 K. The small magnetoresistance for $B \parallel b$ results from the longitudinal configuration $I \parallel B$. (d) SdH oscillations extracted from the data in (c) as a function of inverse field $1/B$. The curves are vertically shifted for clarity. A third polynomial was fitted to each $\rho(B)$ curve to determine a smoothly varying background $\rho_{background}$ and was subtracted to obtain an oscillatory part $\rho_{osc}$. A normalized quantity $\rho_{osc}/\rho_{background}$ is shown. (e) Fourier transforms of the SdH oscillations in the compressed and elongated samples for selected field directions. The field angle $\theta_{ca}$ ($\theta_{cb}$) is that between the $c$ axis and $B$ in the $ca$ ($cb$) plane. The spectra are vertically shifted for clarity. The two samples give consistent results (compare the elongated-sample spectra at $\theta_{ca}$ = 49.5$^{\circ}$ and $\theta_{cb}$ = 100.5$^{\circ}$ with the compressed-sample spectra at $\theta_{ca}$ = 50.5$^{\circ}$ and $B \parallel b$, respectively).}
\end{figure*}
\begin{figure*}
\includegraphics[width=16 cm]{v3fig2.pdf}
\caption{\label{fig1}(color online) (a) Angle dependence of the experimental SdH frequencies in the compressed sample. The mark sizes indicate the oscillation amplitudes. The $ca$-plane (blue squares) and $cb$-plane (red circles) data are superimposed. The solid lines indicate the fundamental frequency branches. The dotted lines indicating the harmonic and combination frequencies are calculated from the solid lines. The dashed line, almost overlapped by a solid line, is a fit to the $\gamma$ branch assuming that the responsible FS pocket is an ellipsoid of revolution. The fit gives $k_F^{ab}$ = 0.050 \AA$^{-1}$ and $k_F^c$ = 0.33 \AA$^{-1}$, $k_F^{ab (c)}$ being the Fermi wave number in the $ab$ plane (along the $c$ axis). No frequency other than 3$\alpha$ is observed above $F$ = 1 kT. (b) Comparison between the calculated and experimental frequencies. For the calculated frequencies, both original and adjusted ones are shown (see text). The solid curves representing the experimental frequencies are the same as those in (a). (c) The FS resulting from the original band structure calculation. It consists of a band-68 hole sheet (left) and band-69 electron sheet (right). Since our band structure calculation does not include spin-orbit coupling, it predicts chains of alternating hole and electron pockets along the grey lines, in accord with Ref.~\onlinecite{Ran09PRB}, but they are so thin that they can not be drawn accurately. The directions $a$, $b$, and $c$ shown in the figure refer to those of the $Fmmm$ orthorhombic unit cell\cite{Rotter08PRB}. (d) The FS in BaFe$_2$As$_2$ determined in the present study. The drawings of the $\alpha$ hole and $\delta$ electron pockets are based on the adjusted band structure calculation, and the $\gamma$ electron pocket is schematically shown, based on the ellipsoid fit in (a). There might be minute FS pockets along the grey lines, though our data show no trace of such pockets. (e) Electronic band structure near $E_F$ along $\Gamma\Delta$FZ. The $\gamma$ pocket is placed at the position marked by the black arrow (see text).}
\end{figure*}
\begingroup
\squeezetable
\begin{table}
\caption{\label{Tab1} $m^*$ and $\tau$. $m_e$ is the free electron mass.}
\begin{ruledtabular}
\begin{tabular}{ccddd}
Fermi & Field & & & \\
surface & direction & \multicolumn{1}{c}{$m^*/m_e$} & \multicolumn{1}{c}{$m^*/m_{band}$} & \multicolumn{1}{c}{$\tau$ (10$^{-12}$ s)}\\
\hline
\multicolumn{1}{c}{$\alpha$ hole} & \multicolumn{1}{c}{$B \parallel a$} & 2.8(1) & 3.0 & 1.4(1)\\
& \multicolumn{1}{c}{$B \parallel b$} & 2.7(1) & 3.6 & 2.8(4)\\
& \multicolumn{1}{c}{$B \parallel c$} & 2.1(1) & 2.8 & 2.3(2)\\
\\
\multicolumn{1}{c}{$\gamma$ electron} & \multicolumn{1}{c}{$B \parallel c$} & 0.9(1) & & 2.6(10)\\
\\
\multicolumn{1}{c}{$\delta$ electron} & \multicolumn{1}{c}{$B \parallel a$} & 1.3(1) & 2.0 & 1.0(2)\\
& \multicolumn{1}{c}{$B \parallel b$} & 2.1(1) & 3.0 & 2.0(4)\\
& \multicolumn{1}{c}{$B \parallel c$} & 2.4(3) & 1.7 & \\
\end{tabular}
\end{ruledtabular}
\end{table}
\endgroup
Figure 1(a) shows $\rho$ as a function of temperature $T$ near the structural/magnetic phase transition for compressed, elongated, and freestanding (= twinned) samples. The $\rho(T)$ curves are consistent with previous data\cite{Chu10Science, Tanatar10PRB}. Figure 1(c) shows the magnetic-field $B$ dependence of $\rho$ in the compressed sample [Fig. 1(b)] for $B \parallel a$, $b$, and $c$ at $T\sim0.17$ K. After subtracting the smooth background, SdH oscillations are clearly visible for all field directions [Fig. 1(d)]. The oscillations continue to below $B$ = 7 T, much lower than previously reported\cite{Sebastian08JPCM, Analytis09PRB}. Figure 1(e) shows Fourier transforms of SdH oscillations in the compressed and elongated samples for selected field directions. The two samples gave consistent results. Figure 2(a) summarizes the angular dependence of the SdH oscillations in the compressed sample. Clear differences are seen between the $ca$ (blue squares) and $cb$ (red circles) planes, confirming appropriate detwinning.
We find three fundamental frequency branches $\alpha$, $\gamma$, and $\delta$ in Fig. 2(a). The first two branches are consistent with (hence named as in) the previous reports\cite{Sebastian08JPCM, Analytis09PRB}, but the last branch $\delta$ has not been observed previously. We attribute the previously reported $\beta$ branch to the second harmonic of $\gamma$ since its angular dependence agrees well with the angular dependence of $\gamma$. This assignment is further supported by the observation that $m^*$ associated with this frequency is twice that of $\gamma$. The $\alpha$ frequency shows only small frequency anisotropy, $F$ for $B \parallel a$ being slightly larger than that for $B \parallel b$, but its amplitude shows clear anisotropy between the $ca$ and $cb$ planes (note the mark sizes in Fig. 2a), which might be explained by anisotropic spin-splitting effects\cite{SM}. The $\gamma$ branch shows no detectable in-plane anisotropy. The angular variation fits that expected for an FS pocket whose shape is a prolate ellipsoid of revolution, as indicated by the dashed line. The angle dependence of the $\delta$ branch indicates that the responsible FS pocket is flattened along the $c$ axis and exhibits the most pronounced frequency anisotropy between the $ca$ and $cb$ planes. The effective mass $m^*$ and electron relaxation time $\tau$ were determined from the $T$- and $B$-dependences of the SdH oscillation amplitudes, respectively, as usual\cite{Shoenberg84}, and are tabulated in Table I.
Our original band structure calculation indicates that bands-68 and -69 cross $E_F$, giving hole and electron sheets of the FS, respectively [Fig. 2(c)]. Each sheet consists of large and small pockets. The frequencies originating from the large hole and electron pockets are in the region of 1 kT, and those from the small pockets are in the region of 0.01 kT for $B \parallel c$ [pale-colored + and x in Fig. 2(b)]. Although the experimental and calculated frequencies differ by a factor of approximately two, it is clear from the angular dependence that the observed $\alpha$ and $\delta$ branches can be assigned to the large hole and electron pockets, respectively, of the calculated FS.
To improve agreement between the experimental and calculated $\alpha$ and $\delta$ frequencies, we shifted the energies of bands-68 and -69 by -0.0032 Ry (-44 meV) and +0.0048 Ry (+65 meV), respectively [Fig. 2(b) thick curves and Fig. 2(d)]. Similar procedures have been used in other multiband metals very successfully\cite{Carrington05PRB, Analytis09PRL}. The agreement now is satisfactory (our criterion for $\alpha$ is the average of the frequencies for the three principal directions), though the observed small angular dependence of $\alpha$ is not reproduced so well as in the original calculation. The adjusted bands-68 and -69 contain 0.0235 holes and 0.0130 electrons per primitive cell, respectively [the primitive cell contains two formula units (fu)]. The densities of states (DOS) at $E_F$ are 6.64 and 6.35 states/Ry per primitive cell for the adjusted bands-68 and -69, respectively. Comparing the measured effective masses with the band masses $m_{band}$ (Table I), we find average mass enhancements $m^*/m_{band}$ of 3.1 and 2.3 for the band-68 $\alpha$ and band-69 $\delta$ pockets, respectively; these are small compared to the enhancements of 3--20 found in KFe$_2$As$_2$\cite{Terashima10JPSJ}. Using these values together with the calculated DOS, we can estimate that bands-68 and -69 contribute 1.8 and 1.2 mJ/K$^2$mol-fu, respectively, to the Sommerfeld coefficient of the specific heat.
We now consider the $\gamma$ pocket. It is reasonable to assume that the $\gamma$ pocket is located at a position where a small pocket appeared in the original calculation. Figure 2(e) shows the relevant part of the band structure; two likely positions are marked by arrows. The band crossings at these points, sometimes referred to as Dirac nodes\cite{Ran09PRB, Richard10PRL}, become anticrossings if spin-orbit coupling, which is absent in our calculation, is included and hence do not necessarily give FS pockets: $E_F$ may occur in a gap. A recent laser ARPES study\cite{Shimojima10PRL}, which is considered bulk sensitive because of the low photon energy of 7 eV, has however shown that the right (anti)crossing marked by the black arrow actually occurs below $E_F$ and produces an electron pocket of non-negligible size. We identify this pocket with our $\gamma$ pocket. The (anti)crossing is situated almost at $E_F$ in the original calculation, and the renormalized Fermi energy for the $\gamma$ pocket is estimated to be 11 meV from the aforementioned ellipsoid fit and measured effective mass. Assuming the mass enhancement to be about three, a band shift of $\sim$30 meV is enough to produce the $\gamma$ pocket. Because of the symmetry, two $\gamma$ pockets occur in the BZ [Fig. 2(d)] and enclose 0.0116 electrons, resulting, within error, in perfect carrier compensation (0.0235 holes vs. 0.0246 electrons). Using the measured effective mass for $\gamma$, we estimate the contribution to the Sommerfeld coefficient to be 1.9 mJ/K$^2$mol-fu. The sum of the contributions from the $\alpha$, $\gamma$, and $\delta$ pockets is 5.0 mJ/K$^2$ mol-fu, which is in excellent agreement with a direct measurement on an annealed single crystal (5.1mJ/K$^2$ mol-fu)\cite{Rotundu10PRB}. The carrier compensation and agreement on the Sommerfeld coefficient rule out existence of unobserved pockets with a comparable volume to the observed ones. Although there might be minute pockets with at least an order-of-magnitude smaller volume along the grey lines in Fig. 2(d), we have seen no trace of such pockets.
Our band structure calculation accounts well for the observed FS if band energy adjustments of at most 65 meV are allowed, despite the fact that the calculated magnetic moment 1.6 $\mu_B$ overestimates the experimental one 0.87 $\mu_B$\cite{Huang08PRL}. The magnitudes of the adjustments are similar to those necessary in a prototypical multiband metal MgB$_2$ (up to $\sim 90$meV)\cite{Carrington05PRB}. Since our calculation is based on the experimental crystal structure and does not include ad hoc procedures such as the As-position adjustment and negative-$U$, those procedures are unjustified. The observed FS is very different from the FS of the DMFT study of Ref.~\onlinecite{Yin11NatPhys}, providing a strong argument against the claim that electronic correlations largely modify the FS of the AFM phase.
The agreement between our FS and those observed in previous ARPES measurements\cite{LiuC09PRL, Yang09PRL, Malaeb09JPSJ, Richard10PRL, Yi11PNAS} is limited. For example, the $\delta$ bright spot, the $\epsilon$ petal electron pocket, and one of the $\alpha$ and $\beta$ hole pockets reported in Ref.~\onlinecite{Yi11PNAS} may correspond to our $\gamma$, $\delta$, and $\alpha$ pockets, respectively, but the other reported pockets are absent in our FS. Further, warped FS cylinders often found in $k_z$-resolved ARPES measurements\cite{LiuC09PRL, Malaeb09JPSJ, Yi11PNAS} are incompatible with our FS composed of closed pockets. It should be noted that the photoelectron escape depth ($\sim$ 5 \AA) for a typical photon energy of 20$\sim$40 eV\cite{Seah79SIA} is shorter than the $c$-axis length (13 \AA), which inevitably limits the $k_z$ resolution.
Our data do not support recent transport studies\cite{Huynh11PRL} attributing observation of the linear-in-$B$ magnetoresistance to Dirac fermion transport in the quantum limit. Our $\rho(B)$ curve for $B \parallel c$ is clearly superlinear [Fig. 1(c)]. The $\gamma$ pocket, a supposed Dirac pocket, does not reach the quantum limit until $B$ becomes comparable to $F_{\gamma}$ = 90 T. The conductivities of the $\alpha$ and $\gamma$ pockets are estimated from $m^*$ and $\tau$ ($B \parallel c$) in Table 1 to be 3.6(2) and 5(2) $\times 10^6$ m$^{-1}$$\Omega^{-1}$, respectively. The sum is already comparable to the measured conductivity, 11(2) $\times 10^6$ m$^{-1}$$\Omega^{-1}$. It is therefore also unlikely that unobserved Dirac pockets, if any, dominate the conduction.
The origin of the in-plane resistivity anisotropy, which increases up to $\rho_b/\rho_a \sim$2 with only a few \% Co-doping\cite{Chu10Science, Tanatar10PRB}, is left elusive. The in-plane mass anisotropy is negligible for the $\alpha$ pocket and opposite to the resistivity anisotropy for the $\delta$ pocket, i.e., $m^*$ along the $b$ axis being smaller than $m^*$ along the $a$ axis [Table 1, note $m^*$ for $B \parallel a (b)$ is associated with cyclotron motion in the $bc (ac)$ plane]. Intriguingly, this does not seem compatible with an optical study\cite{Dusza11EPL}, which indicates that the Drude spectral weight ($\omega_p^2=ne^2/\epsilon_0 m$ in a free-electron model) for $E \parallel a$ increases and becomes much larger than that for $E \parallel b$ below $T_N$. It might be better reconciled with a more recent study\cite{Nakajima11PNAS} suggesting an isotropic Drude component.
\begin{acknowledgments}
TT thanks Takahiro Shimojima, Tamio Oguchi, and Hiroaki Ikeda for helpful discussions.
\end{acknowledgments}
|
\section{INTRODUCTION}
The gas disk surrounding a young stellar object is usually argued to be a turbulent accretion disk when the star is still in the process of accumulating materials \citep[see, e.g.,][]{FKR02}. While the molecular viscosity of the gas is negligibly small, shear stress resulting from turbulence can effectively drive angular momentum transport in the disk. At the same time, the dynamics of protoplanetary objects embedded in the disk is significantly affected by their interactions with the turbulent gas, at a stage when these objects are still amassing solids to grow in size \citep[e.g.,][and references therein]{PT06}. Understanding particle dynamics in a turbulent gas disk, therefore, will provide insight into the paths along which new planets may eventually form.
One of the most promising mechanisms to drive the turbulence is through the magneto-rotational instability \cite[MRI; see, e.g.,][and references therein]{BH98}. A weakly magnetized, differentially rotating gas disk is unstable to linear perturbations, and after nonlinear effects set in, the gas presumably reaches a statistically steady, sustained, turbulent state. The underlying processes that determine the properties of this saturated state are still under active research \citep[e.g.,][]{JJK08,LLB09,eV09,mP10,OM11}. Nevertheless, it is known that the strength of this magneto-rotational turbulence depends on the net magnetic flux threading the disk \citep[e.g.,][hereafter YMM09]{HGB95,JKH06,YMM09}, and thus can be controlled to the desired level and treated as a parameter.
The gravitational torques exerted on an embedded solid object due to density fluctuations in magneto-rotational turbulence have been shown to be stochastic, rendering the motion of the object a random walk (\citealt{LSA04,NP04,N05,OMM07}; YMM09; \citealt{NG10}) on top of the ordered migration expected in a laminar disk \citep{JGM06,AB09,KAB11}. Note that the ordered migration may be either inward or outward in different regions of the disk \citep{PB10,PB11,LPM10}. In addition to inducing this radial diffusive migration, the same gravitational force drives the diffusion of orbital eccentricity of the solid objects, increasing their velocity dispersion (\citealt{OIM07,IGM08}; YMM09; \citealt{NG10}). The effect on the orbital inclination, however, has not been investigated yet, as that requires the inclusion of the vertical component of the gravity from the host star. Overall, the gravitational influence of the turbulent gas on particle orbital properties may play an important role in the course of planet formation.
To assess the significance of this process in planet formation scenarios, numerical simulations simultaneously evolving magneto-rotational turbulence and particle dynamics can be conducted and the resulting particle orbital evolution can be statistically measured. However, a consistent picture of the actual magnitude of this effect has been elusive. Using the local-shearing-box simulations, we reported in YMM09 that the response of test particles to the gravity of the turbulent gas is systematically lower than what was previously reported in global disk models. We also found in the same study that the effect significantly depends on the horizontal box size of the numerical models. Recently, \citet{NG10} have found similar behavior in their local models and argued that large box size might be necessary to see convergence in the stochastic torque and forcing generated by the turbulence.
Extending the local unstratified disk models of YMM09, we now consider disks with vertical density stratification by including the linearized vertical gravity from the central star. We examine whether a stratified disk model gives results consistent with a comparable unstratified model and whether convergence in particle orbital evolution, if any, can be achieved with local-shearing-box simulations. In Section~\ref{S:model}, we describe in detail our numerical models. We report the properties of our modeled magneto-rotational turbulence in Section~\ref{S:turbulence} and the statistical evolution of particle orbital properties in Section~\ref{S:particles}. We analyze the gravitational field generated by the turbulence in Section~\ref{S:force} and discuss the convergence issues of our local models in Section~\ref{S:box_size}. We conclude our discussion in Section~\ref{S:summary}.
\section{NUMERICAL MODELING} \label{S:model}
Directly extending the work we reported in YMM09, we continue to use the Pencil Code\footnote{The Pencil Code is publicly available at \texttt{http://code.google.com/p/pencil-code/}.} to model particles moving in magneto-rotational turbulence. We describe in detail the equations and the relevant numerical techniques we implemented in the code.
\subsection{Magnetohydrodynamics} \label{SS:mhd}
We use the local shearing box approximation \citep[e.g.,][]{GL65,BN95,HGB95}. A local shearing box is a small Cartesian box at a large distance $R$ from the host star such that the center of the box revolves at the Keplerian angular speed $\Omega_K$. The box is always oriented with the $x$-axis directed radially and the $y$-axis azimuthally. In contrast to YMM09, we include the linearized vertical gravity from the host star so that the disk is vertically stratified. We again impose a vertical, external magnetic field $\vec{B}_\mathrm{ext} = B_\mathrm{ext}\hat{\vec{z}}$ to maintain a non-zero magnetic flux. The MHD equations then become
\begin{eqnarray} \label{E:mhd}
\partial_t\rho
- \frac{3}{2}\Omega_K x\partial_y\rho
+ \nabla\cdot(\rho\vec{u})
&=& f_D,\label{E:mhd_cont}\\
\partial_t\vec{u}
- \frac{3}{2}\Omega_K x\partial_y\vec{u}
+ \vec{u}\cdot\nabla\vec{u}
&=& -\frac{1}{\rho}\nabla p
+ \left(2\Omega_K u_y\hat{\vec{x}}
- \frac{1}{2}\Omega_K u_x\hat{\vec{y}}
- \Omega_K^2 z\hat{\vec{z}}\right)\nonumber\\
&&+ \frac{1}{\rho}\vec{J}\times\left(\vec{B} +
\vec{B}_\mathrm{ext}\right)
+ \vec{f}_V,\label{E:mhd_mom}\\
\partial_t\vec{A}
- \frac{3}{2}\Omega_K x\partial_y\vec{A}
&=& \frac{3}{2}\Omega_K A_y\hat{\vec{x}}
+ \vec{u}\times\left(\vec{B} + \vec{B}_\mathrm{ext}\right)
+ \vec{f}_R,\label{E:mhd_ind}
\end{eqnarray}
where $\rho$ is the gas density, $\vec{u}$ is the gas velocity relative to the background shear flow, $p$ is the gas pressure, $\vec{J} = \nabla\times\vec{B} / \mu_0$ is the electric current density, $\vec{B} = \nabla\times\vec{A}$, and $\mu_0$ is the vacuum permeability. The terms $f_D$, $\vec{f}_V$, and $\vec{f}_R$ are numerical dissipation terms, including both hyper-diffusion and shock diffusion, that are needed to stabilize the scheme. The reader is referred to YMM09 for their description.
To close the system of equations~\eqref{E:mhd_cont}--\eqref{E:mhd_ind}, we assume an isothermal ``equation of state'', $p = c_s^2\rho$, where $c_s$ is the isothermal speed of sound. This is an ideal limit of the energy equation, in which heat diffusion is effectively instantaneous. Gas buoyancy should thus be absent and the vertical mixing in the resulting magneto-rotational turbulence might be artificially enhanced \citep{BG09}. Nevertheless, \citet{OM09} showed that vertical motions are not completely suppressed even in the case of adiabatic gas, where no heat diffusion occurs, and the differences in the mean kinetic and the magnetic energies of a vertically stratified gas disk between the isothermal and the adiabatic limits may only amount to about a factor of two.
We set up the gas density so that the gas is in vertical hydrostatic equilibrium initially:
\begin{equation} \label{E:rho_0}
\rho_0(z) = \rho_m\exp\left(-\frac{z^2}{H^2}\right),
\end{equation}
where $\rho_m$ is the mid-plane gas density and $H = \sqrt{2}c_s / \Omega_K$ is the vertical disk scale height.\footnote{Note that there exists a factor of $\sqrt{2}$ difference between our definition of disk scale height $H$ (Equation~\eqref{E:rho_0}) and some other authors' in the literature.\label{scale_height}} We fix the vertical dimension of the computational domain at $|z| \leq 2H$. The initial magnetic vector potential $\vec{A}$ is set to zero, while the external magnetic field is fixed at such a level that the corresponding plasma $\beta_\mathrm{ext}(z) \equiv 2\mu_0 c_s^2\rho_0 / B_\mathrm{ext}^2$ is $\beta_\mathrm{ext}(0) = 6.2\times10^3$ in the mid-plane. Gaussian noise of magnitude $10^{-3}\,H/P$ in each component of the gas velocity $\vec{u}$ is generated to seed the MRI, where $P = 2\pi / \Omega_K$ is the orbital period at the center of the shearing box.
Our assumption of a net vertical magnetic field allows us to compare our models directly to the unstratified models presented in YMM09. The strength of this field controls the strength of the MRI-driven turbulence, so imposing it is an effective way to adjust the viscous stress generated by the MRI. A similar procedure has been used in other recent publications including \citet{JO07,JYK09}, \citet{NG10}, \citet{SI09}, and \citet{SMI10}. Some justification for this assumption can be found in the argument that at high altitude where the gas is tenuous, the magneto-convective motions tend to turn horizontal magnetic fields into vertical ones \citep[e.g.,][]{PW82,HT88,BN95}. Therefore, we take the limit that $B_x = B_y = 0$ at the $z$-boundary while allowing the vertical component ($B_z$) to change freely. In terms of the vector potential $\vec{A}$, this boundary condition translates to $\partial_z A_x = \partial_z A_y = A_z = 0$. In the horizontal directions, we use the standard sheared periodic boundary conditions for the magnetic field, i.e., the field is periodic after a shift compensating for the Keplerian shear flow (\citealt{HGB95}; see also Section~\ref{SS:poisson}).
We explored several different vertical boundary conditions for the gas and the field. We found that outflow boundary conditions for both led to numerical instability during the initial growth of the MRI. Although \citet{SI09} and \citet{SMI10} did publish models using outflow boundary conditions, their disk lost significant amounts of mass through the boundary during their run. We also found that periodic boundary conditions on both gas and field led to random crashes after tens of orbits of saturated turbulence for reasons that remain unclear. On the other hand, the combination of the vertical field boundary condition with periodic boundary conditions in the gas leads to a stable numerical result for the hundreds of orbits required for us to accurately measure the evolution of particle orbital properties \citep[see also][]{PW82}. The horizontal boundary conditions are again sheared periodic. In this setup, the total mass is conserved while saturated turbulent motion can be maintained near the vertical boundaries at a steady level.
While it is interesting in itself to study the coronal structure and possibly the net flow of the gas \citep{SI09,SMI10,FD11}, we specifically exclude the modeling of the coronal regions to maintain the steadiness of the magneto-rotational turbulence and thus improve the statistical measurement of the particle orbital properties. The gravitational acceleration due to the host star across the boundary does not present any numerical difficulty. Even though the vertical component of the gravitational acceleration changes sign when a fluid element is wrapped around the vertical boundary, the vertical velocity component of the element retains sign such that its velocity and acceleration always have the same sign after reemerging from the opposite boundary, resembling material ballistically falling back down onto the disk. Numerically, no z-derivative of the gravitational acceleration due to the host star is ever evaluated (see Equation~\eqref{E:mhd_mom}), so the discontinuity of the acceleration across the vertical boundary is not an issue here. Turbulence properties near the disk mid-plane, where the gravitational influence of the turbulent gas on the particles is strongest \citep{OMM07}, appear insensitive to the choice of vertical boundary conditions \citep{JYK09,DSP10,GG11}, as we also demonstrate in Section~\ref{SS:bce}.
For large box simulations, concerns have been raised that artificial numerical diffusion varying with $x$ may occur due to the radial dependence of the shear velocity and thus corresponding truncation errors on a fixed grid \citep{JGG08}. Explicit treatment of the shear advection may be needed to eliminate these truncation errors. In the Pencil Code, such a technique, dubbed shear advection by Fourier interpolation (SAFI), has been implemented \citep{JYK09}. In our work, we implement SAFI for all of our simulations to remove this undesirable numerical effect.
\subsection{Poisson Equation with Isolated Boundary Condition in $z$-direction} \label{SS:poisson}
To find the gravitational influence of the turbulent gas on the movement of solid particles, we need to solve the Poisson equation for the fluctuation potential $\Phi_1$:
\begin{equation} \label{E:poisson}
\nabla^2\Phi_1 = 4\pi G\rho_1,
\end{equation}
where $G$ is the gravitational constant and $\rho_1(x,y,z) \equiv \rho(x,y,z) - \rho_0(z)$ is the density fluctuation with respect to the basic state of the gas stratification $\rho_0(z)$, which is given by Equation~\eqref{E:rho_0}. Note that we have neglected the self-gravity of the gas in Equation~\eqref{E:mhd_mom} on the assumption that the disk is gravitationally stable (see YMM09), so the solution to Equation~\eqref{E:poisson} does not affect the gas dynamics.
Although the vertical boundary conditions for the gas density are assumed to be periodic in the MHD equations (see Section~\ref{SS:mhd}), we find it inappropriate to assume the same in solving the Poisson equation~\eqref{E:poisson}. Consider, for example, density perturbations near the top edge of the computational domain. If periodic boundary conditions in the vertical direction were adopted, particles near the mid-plane would experience the gravity of the same density perturbations at almost equal distance from above and below, resulting in cancellation of the gravitational force exerted by these perturbations on the particles. This would amount to undesirable, potentially large, numerical errors for calculating the gravity of density perturbations at high altitude.
Therefore, we instead adopt isolated or open boundary conditions in the vertical direction to solve Equation~\eqref{E:poisson}. With these boundary conditions, we assume that there are no fluctuations in density outside the vertical boundary, i.e., $\rho_1(x,y,z) = 0$ for $|z| > 2H$. The boundary conditions in the horizontal directions remain sheared periodic.
To implement approximate isolated vertical boundary conditions with minimal computational cost, we use a variation of the \citet{HE88} fast algorithm to solve the Poisson equation~\eqref{E:poisson}, combining it with the Fourier interpolation technique to achieve sheared periodicity \citep{BI07,JO07}. The essence of this algorithm is to double the computational domain in the $z$-direction by appending regions with zero density fluctuation\footnote{In our case, we append $\rho_1(x,y,z) = 0$ for $2H < z < 6H$ so the vertical domain now covers $-2H < z < 6H$.} and apply fast Fourier transforms to this expanded domain. Although the vertical boundary conditions remain periodic with Fourier transforms, the $1/r^2$ nature of gravity means that influence from the periodic copies, now at least $2 L_z$ away, of the original computational domain of size $L_z$ onto itself is significantly reduced.
We now describe the algorithm to solve Equation~\eqref{E:poisson} in detail. After expanding the vertical dimension of the computational domain and initializing the expanded region with zero density fluctuations, $\rho_1(x,y,z)$ is Fourier transformed in the $y$-direction into $\hat{\rho}_1(x,k_y,z)$, where $k_y$ is the $y$-wavenumber. The result is phase shifted to recover periodicity in $x$-direction: $\check{\rho}_1(x,k_y,z) = \hat{\rho}_1(x,k_y,z)\exp\left[-i(3/2)\Omega_K k_y x\delta t \right]$, where $\delta t$ is the time step. Then $\check{\rho}_1(x,k_y,z)$ is Fourier transformed in the $x$-direction into $\tilde{\rho}_1(k_x,k_y,z)$, where $k_x$ is the $x$-wavenumber. For the modes $k_x^2 + k_y^2 > 0$, $\tilde{\rho}_1(k_x,k_y,z)$ is Fourier transformed in $z$-direction to find $\breve{\rho}_1(k_x,k_y,k_z)$, where $k_z$ is the $z$-wavenumber. The fluctuation potential in Fourier space $\breve{\Phi}_1(k_x,k_y,k_z)$ is calculated using the usual convolution theorem:
\begin{equation} \label{E:convol}
\breve{\Phi}_1(k_x,k_y,k_z)
= -\frac{4\pi G}{\left(k_x + 3\Omega_K k_y\delta t / 2\right)^2 +
k_y^2 + k_z^2}
\breve{\rho}_1(k_x,k_y,k_z),\quad
\textrm{for}\ k_x^2 + k_y^2 > 0.
\end{equation}
It is then inverse Fourier transformed in $k_z$ to find $\tilde{\Phi}_1(k_x,k_y,z)$.
In principle, the mode $k_x = k_y = 0$ can similarly be treated by Fourier transforms in the vertical direction and convolved by Equation~\eqref{E:convol} as above. However, this mode has an analytical solution that we can employ with marginal computational cost to improve the solution to Equation~\eqref{E:poisson}.\footnote{We note that use of Equation~\eqref{E:convol} would yield a 50\% relative error in the vertical acceleration due to a horizontal uniform slab at a distance $L_z / 2$ away (after domain doubling) when compared to the exact acceleration derived from Equation~\eqref{E:sheet}.} It represents a horizontally infinite, vertically thin mass layer of constant density at a given altitude $z$, where the density is equal to the horizontal average of the density fluctuation at $z$. (The layer can be considered as either a slab of finite size $\Delta z$ and constant volume density $\tilde{\rho}_1(k_x=0,k_y=0,z)$ or as an infinitely thin sheet of surface density $\tilde{\rho}_1(k_x=0,k_y=0,z) \Delta z$; the solutions outside the layer are the same by symmetry and Gauss' law.) The gravitational acceleration due to this mode is constant above and below $z$. Therefore, the resultant fluctuation potential can be calculated by
\begin{equation} \label{E:sheet}
\tilde{\Phi}_1(k_x=0,k_y=0,z) = 2\pi G\int\,\tilde{\rho}_1(k_x=0,k_y=0,z')|z - z'|dz',
\end{equation}
where we have arbitrarily defined the reference potential to be zero at each altitude. The discretized version of Equation~\eqref{E:sheet} reads
\begin{equation} \label{E:dis_sheet}
\tilde{\Phi}_1(k_x=0,k_y=0,z_j) =
2\pi G\sum_k \tilde{\rho}_1(k_x=0,k_y=0,z_k)|j - k|\Delta z^2.
\end{equation}
Equation~\eqref{E:dis_sheet} remains exact for this mode given that we do not consider reconstruction of the density field within each cell. Finally, we reverse the process to inverse Fourier transform $\tilde{\Phi}_1(k_x,k_y,z)$ in $k_x$ and $k_y$ to derive the fluctuation potential in real space $\Phi_1(x,y,z)$, incorporating the corresponding phase shifts with opposite sign.
In practice, it is not necessary to allocate storage space for the whole extended domain in the $z$-direction. The appended zero density is only involved in the calculation between the forward and inverse Fourier transforms in $x$ and $y$. In addition, the convolutions in $z$ (Equations~\eqref{E:convol} and~\eqref{E:dis_sheet}) for different modes $(k_x,k_y)$ are independent of each other. Therefore, these convolutions can be distributed among processors and performed sequentially by allocating only a single one-dimensional working array along the $z$-direction.
\subsection{Particle Dynamics} \label{SS:pd}
We continue to use the zero-mass approximation to model our solid particles as in YMM09. In this approximation, particles behaves as test particles and only respond to the gravity of the host star and the gas. We ignore the viscous drag forces between particles and gas. This remains a good approximation for kilometer-sized planetesimals \citep[e.g.,][]{OMM07}. We also ignore mutual gravitational interactions between particles, which helps us isolate the net effect induced by hydromagnetic turbulence.
The equations of motion for each particle, therefore, become
\begin{mathletters} \label{E:eomp}
\begin{eqnarray}
\frac{d\mathbf{x}_p}{dt}
&=& \mathbf{u}_p
- \frac{3}{2}\Omega_K x_p\hat{\mathbf{y}},\label{E:eomp1}\\
\frac{d\mathbf{u}_p}{dt}
&=& \left(2\Omega_K u_{p,y}\hat{\mathbf{x}}
- \frac{1}{2}\Omega_K u_{p,x}\hat{\mathbf{y}}
- \Omega_K^2 z_p\hat{\mathbf{z}}\right)
+ \mathbf{g}_0 - \nabla\Phi_1.\label{E:eomp2}
\end{eqnarray}
\end{mathletters}
The vector $\mathbf{x}_p = \left(x_p, y_p, z_p\right)$ is the position of the
particle in the shearing box, while $\mathbf{u}_p = \left(u_{p,x}, u_{p,y}, u_{p,z}\right)$ is the velocity of the particle relative to the background shear flow. In Equation~\eqref{E:eomp2}, the terms inside the parentheses stem from the linearized gravity of the host star and the Coriolis and centrifugal forces in the co-rotating frame of the shearing box. The remaining terms on the right-hand side represent the acceleration attributed to the gas: $\mathbf{g}_0$ is the gravitational acceleration due to the basic state of the gas stratification and has the analytical expression
\begin{equation} \label{E:g_0}
\mathbf{g}_0(z) = -2\pi^{3/2}G\rho_m H
\,\mathrm{erf}\left(\frac{z}{H}\right)
\hat{\mathbf{z}},
\end{equation}
while $\Phi_1$ is the gravitational potential due to density fluctuations with respect to the basic state and is the solution of the Poisson Equation~\eqref{E:poisson} (Section~\ref{SS:poisson}). Note that by separating the density perturbation from the equilibrium density stratification and treating the gravity of the equilibrium state exactly, we improve the accuracy of the gravitational potential resulting from the turbulence.
The position $\mathbf{x}_p$ and velocity $\mathbf{u}_p$ of each particle is evolved in time by integrating the equations of motion~\eqref{E:eomp} simultaneously with the third-order Runge-Kutta steps advancing the MHD equations. In addition to the Courant conditions set by the MHD equations, the time-step is limited by the absolute maximum of velocity $\vec{u}_p$ so that no particles can cross more than half a grid zone each timestep. We calculate the gradient of the fluctuation potential $\nabla\Phi_1$ on the grid and then quadratically interpolate it onto each particle.
We uniformly distribute $128^2$ particles in a horizontal plane. We do not allow these particles to move until after 20 orbital periods, when the turbulence has reached a statistically steady state. The particles can have an initial eccentricity $e_0$ by setting an initial velocity $\vec{u}_p$ so that they are at the apogee of their orbits (see YMM09). They can also have an initial inclination $i_0$ by placing the initial particle plane at a distance to the mid-plane. Particles are wrapped around when crossing any of the six boundary planes of the shearing box.
\section{PROPERTIES OF MAGNETO-ROTATIONAL TURBULENCE} \label{S:turbulence}
In this section, we present several properties of the magneto-rotational turbulence in our numerical models and discuss their convergence with the box size and the resolution.
Figure~\ref{F:mri_t} shows the horizontally averaged density fluctuation, inverse plasma $\beta$, and $\alpha$ parameter in the mid-plane ($z = 0$) as a function of time, for resolutions up to 64~points per $H$ and horizontal box dimensions up to $L_x = L_y = 16~H$. The magnitude of the density fluctuation is indicated by the rms value relative to the equilibrium density $\langle\rho_1^2\rangle^{1/2} / \rho_0$, the plasma $\beta \equiv 2\mu_0\rho_0 c_s^2 / \langle\left|\vec{B} + \vec{B}_\mathrm{ext}\right|^2\rangle$ is the ratio of the equilibrium thermal pressure to the magnetic pressure, and $\alpha$ is calculated by \citep[e.g.,][]{aB98}
\begin{equation}
\alpha = \frac{\sqrt{2}}{3}
\frac{\langle\rho u_x u_y - B_x B_y / \mu_0\rangle}{\rho_0 c_s^2},
\end{equation}
which includes both Reynolds and magnetic stresses.\footnote{The bracket $\langle\ \rangle$ denotes the horizontal average of the enclosed variable at any given altitude $z$ and thus is a function of $z$.}
\placefigure{F:mri_t}
The magneto-rotational turbulence in the mid-plane of our models reaches a statistically steady state at about $t = 20P$. In this saturated state, we see in general that the larger the horizontal size of the box, the smaller the amplitude of oscillation in the turbulence properties. Little difference exists between an 8$\times$8$\times$4$H$ box and a 16$\times$16$\times$4$H$ box at the same resolution, indicating convergence with horizontal box dimension. On the other hand, the amplitude of oscillation at saturation level increases with resolution for fixed box dimensions. There exists some evidence that the magnetic activity and thus the $\alpha$ value increase with resolution, especially for the case of 2$\times$2$\times$4$H$ box, but this effect seems to become smaller with larger box sizes. We also note that there exists a curious trend of increase in density perturbation with time for the case of an 8$\times$8$\times$4$H$ box at the resolution of 64~points/$H$. In spite of these trends, the time average of each turbulence properties remains roughly the same for different resolutions and thus shows convergence. We note that the turbulence properties at the saturated state in the mid-plane of our vertically stratified disks agrees with those in our unstratified disks of YMM09 \citep[see also][]{GG11}. The means and standard deviations of these properties for each set of resolution and box dimensions are reported in Table~\ref{T:mri}, which includes the magnetic stress and the Reynolds stress as well.
\placetable{T:mri}
Figure~\ref{F:mri_z} shows the time-averaged vertical profiles of density perturbation, inverse plasma $\beta$, and $\alpha$ parameter at saturation level over a period of $100P$. Each of the three properties increases with $|z|$, indicating stronger turbulence at higher altitude. The increasing activity with altitude is related to the increasing importance of the external magnetic field, which can be quantified by $\beta_\mathrm{ext}(z)$ (Section~\ref{SS:mhd}). This agrees with the trend found in models of unstratified disks (\citealt{HGB95,JKH06}; YMM09). Notice that $\beta_\mathrm{ext}(\pm2H) = 1.1\times10^2 \gg 1$ at our highest altitude, and thus the corona is not modeled in our simulations. Given that our vertical height only amounts to 2$H$, the gas flow in our computational domain remains marginally stable against magnetic buoyancy during all stages of the development of the MRI \citep[e.g.,][]{SKH10,GG11,SHB11}. As shown in Figure~\ref{F:mri_z}, the time-averaged inverse plasma beta near the vertical boundary amounts to only about 0.4, and thus on average, thermal pressure still dominates magnetic pressure there.
\placefigure{F:mri_z}
\section{ORBITAL PROPERTIES OF MASSLESS PARTICLES} \label{S:particles}
We now report the response of zero-mass particles to the gravity of the density fluctuations of the statistically steady, numerically convergent magneto-rotational turbulence described in Section~\ref{S:turbulence}. The reference time $t = 0$ in the following discussion is the time at which the turbulent gas reaches its saturated state and the particles are allowed to move.\footnote{We choose this reference time to be $t = 20P$ as reported in Section~\ref{S:turbulence}. See Figure~\ref{F:mri_t}.}
\subsection{Mean Orbital Radius} \label{SS:radius}
The evolution of the mean orbital radius of one particle can be found by averaging the radial position $x$ of the particle over each epicycle motion. For the case of ideal unstratified disks, the distribution of particles in terms of the orbital radius change can be described by a time-dependent normal distribution centered at zero:
\begin{equation} \label{E:dist_dx}
f(\Delta x, t) = \frac{1}{\sqrt{2\pi}\sigma_x(t)}
\exp\left[-\frac{\Delta x^2}{2\sigma_x^2(t)}\right],
\end{equation}
where $\Delta x$ is the orbital radius change from the initial radius $x_0$ at $t = 0$ and $\sigma_x(t)$ is the time-dependent standard deviation. We reported in YMM09 that $\sigma_x(t)$ depends on the properties of the gas disk and can be concisely expressed by
\begin{equation} \label{E:sigma_x}
\sigma_x(t) = C_x\xi H\left(\frac{t}{P}\right)^{1/2},
\end{equation}
where $C_x$ is a dimensionless proportionality constant\footnote{$C_x$ as well as other dimensionless constants introduced in the following discussions may depend on the $\alpha$ parameter. This dependency is not investigated in this work.} and $\xi \equiv 4\pi G\rho_0 P^2$ is a dimensionless quantity indicating the strength of the gas gravity. Equations~\eqref{E:dist_dx} and \eqref{E:sigma_x} demonstrate the diffusive nature of particle radial migration driven by magneto-rotational turbulence.
For the case of ideal stratified disks studied here, we again find the evolution of particle orbital radius can be described by Equations~\eqref{E:dist_dx} and \eqref{E:sigma_x}. The dimensionless quantity $\xi$ is now defined by
\begin{equation}
\xi \equiv 4\pi G\rho_m P^2 = 4(2\pi)^{3/2} / Q_g
\end{equation}
in terms of the mid-plane gas density $\rho_m$, where $Q_g$ is the Toomre stability parameter for the gas. In Table~\ref{T:orbit}, we report the measured values of the constant $C_x$ from our simulations at different resolutions and horizontal box sizes for the case of particles with zero initial inclination. Comparing the value from the 2$\times$2$\times$4$H$ stratified model at a resolution of 64~points/$H$ with that from the unstratified model at the same resolution and horizontal box size (see Equation~(15) of YMM09), we find excellent agreement in the two $C_x$ values.
As can be seen from Table~\ref{T:orbit} and Figure~\ref{F:orbit}, the value of $C_x$ remains roughly the same with different resolutions for fixed box dimensions (except the anomaly shown by the 2$\times$2$\times$4$H$ box at a resolution of 64~points/$H$). Conversely, it is significantly dependent on the horizontal box size $L_h \equiv L_x = L_y$. The larger the size, the stronger the influence of the turbulence on the particle orbital radius. This relationship can be represented by the following power-law fit:
\begin{equation} \label{E:cx}
C_x \simeq 6.6\times10^{-5}\,(L_h / H)^{1.35}.
\end{equation}
We find no evidence of convergence with horizontal box size up to $L_h = 16H$, the largest size we have investigated.
\placefigure{F:orbit}
Equation~\eqref{E:sigma_x} can be transformed into the diffusion coefficient $D(J)$ introduced by \citet{JGM06} for describing the radial random walks of orbiting particles induced by turbulent torques, in terms of the Keplerian orbital angular momentum $J$. The reader is referred to YMM09 for a detailed description of this procedure. We emphasize here that this transformation involves no assumption about the correlation time of the stochastic torques and is thus a direct measurement of $D(J)$. Using a heuristic choice of dimension for the diffusion coefficient, \citet{JGM06} defined a dimensionless parameter $\epsilon$ to represent the magnitude of $D(J)$. We report our measured values of $\epsilon$ in Table~\ref{T:orbit}. The dependence of $\epsilon$ on the horizontal box size $L_h$ in our models for $L_h \gtrsim 4H$ is shown in Figure~\ref{F:epsilon_gamma} and can be written as
\begin{equation} \label{E:epsilon}
\epsilon \simeq 6.5\times10^{-6}\,(L_h / H)^{2.69}.
\end{equation}
\placefigure{F:epsilon_gamma}
\subsection{Eccentricity} \label{SS:ecc}
The amplitude of each epicyclic oscillation of a particle gives the instantaneous eccentricity of the particle orbit. We reported in YMM09 that in ideal unstratified disks, the distribution of particles in terms of the eccentricity deviation should be a time-dependent normal distribution centered at zero, as long as the particles have non-negligible initial eccentricity (c.f.\ Equation~\eqref{E:dist_dx}):
\begin{equation} \label{E:dist_de}
f(\Delta e, t) = \frac{1}{\sqrt{2\pi}\sigma_e(t)}
\exp\left[-\frac{\Delta e^2}{2\sigma_e^2(t)}\right],
\end{equation}
where $\Delta e$ is the eccentricity deviation from the initial eccentricity $e_0$ at $t = 0$ and the time-dependent standard deviation $\sigma_e(t)$ can be written as
\begin{equation} \label{E:sigma_e}
\sigma_e(t) = C_e\xi\left(\frac{H}{R}\right)\left(\frac{t}{P}\right)^{1/2},
\end{equation}
where $C_e$ is a dimensionless proportionality constant. We emphasize that the time-dependent Rayleigh distribution\footnote{Note that the standard deviation of a Rayleigh distribution is equal to $\sigma_e\sqrt{(4-\pi)/2}$.} found for particles with zero (or negligible) initial eccentricity
\begin{equation} \label{E:dist_e}
f(e,t) = \frac{e}{\sigma_e^2(t)}\exp\left[-\frac{e^2}{2\sigma_e^2(t)}\right]
\end{equation}
is a manifestation of Equation~\eqref{E:dist_de} since the eccentricity $e$ is a positive definite quantity. Equations~\eqref{E:dist_de} and \eqref{E:dist_e} share the same time-dependent parameter $\sigma_e(t)$, and there exists no evidence that $\sigma_e(t)$ depends on the initial eccentricity $e_0$.
We find that the same evolution of particle distribution in eccentricity deviation also holds for the case of ideal stratified disks. The measured values of the constant $C_e$ for different resolutions and horizontal box sizes when particles have zero initial inclination are listed in Table~\ref{T:orbit}. The same comparison performed in Section~\ref{SS:radius} for the orbital radius evolution indicates that the eccentricity evolution in a stratified disk again agrees very well with that in an unstratified disk (see Equation~(17) of YMM09). These comparisons demonstrate that a local unstratified disk model gives results consistent with the mid-plane of a local stratified disk model that has the same physical conditions except vertical stratification.
As in the case of orbital radius discussed in Section~\ref{SS:radius}, the eccentricity evolution of the particles does not noticeably depend on the resolution for given box dimensions, while it is strongly affected by the horizontal box size (Table~\ref{T:orbit} and Figure~\ref{F:orbit}). A power-law regression gives
\begin{equation} \label{E:ce}
C_e \simeq 7.2\times10^{-5}\,(L_h / H)^{1.08},
\end{equation}
which is close to a linear relation to the horizontal box size. We discuss the box-size effect on both the orbital radius and the eccentricity in Section~\ref{S:box_size}.
The eccentricity driven by magneto-rotational turbulence enhances orbital crossing among planetesimals and thus increases the chance of collisions between them. \citet{IGM08} defined a dimensionless parameter $\gamma$ to represent the strength of this effect. Equation~\eqref{E:sigma_e} can be used to estimate the value of $\gamma$, and the reader is referred to YMM09 for a detailed description of this procedure. We report in Table~\ref{T:orbit} our measured values of $\gamma$. As shown in Figure~\ref{F:epsilon_gamma}, the dependence of $\gamma$ on the horizontal box size $L_h$ in our models for $L_h \gtrsim 4H$ can be described by
\begin{equation} \label{E:gamma}
\gamma \simeq 3.6\times10^{-5}\,(L_h / H)^{1.08}.
\end{equation}
\subsection{Inclination}
The only orbital property of a particle that cannot be measured in an unstratified disk model is the inclination $i$. In a stratified disk, vertical linear gravity from the host star (the third term in the parentheses on the right-hand side of Equation~\eqref{E:eomp2}) provides a restoring force such that particles oscillate in the $z$-direction about the mid-plane, as demonstrated in Figure~\ref{F:zpt1}. Note that because the particles also experience the gravity of the gas, the period of the oscillation is $P\left(1 + \xi / 4\pi^2\right)^{-1/2}$ in the linear limit (derived from Equations~\eqref{E:eomp2} and \eqref{E:g_0}), which is slightly shorter than the orbital period $P$. We can calculate the induced inclination (in radians) for a single particle by $i \approx \left(z_\mathrm{max} - z_\mathrm{min}\right) / 2R$, in which $z_\mathrm{max}$ and $z_\mathrm{min}$ are the maximum and the minimum vertical positions in one oscillation, respectively, provided that $z / R \ll 1$.
\placefigure{F:zpt1}
Figure~\ref{F:inc}(a) shows the histograms of particles with zero initial inclination sorted in bins of instantaneous inclination at three different times. Similar to the eccentricity distribution for particles with $e_0 = 0$ described by Equation~\eqref{E:dist_e}, the inclination distribution resembles a time-dependent Rayleigh distribution
\begin{equation} \label{E:dist_i}
f(i,t) = \frac{i}{\sigma_i^2(t)}\exp\left[-\frac{i^2}{2\sigma_i^2(t)}\right],
\end{equation}
where $\sigma_i(t)$ is a time-dependent parameter denoting the width of the distribution.
\placefigure{F:inc}
One might expect the inclination distribution of particles with nonzero initial inclination would evolve as a time-dependent normal distribution with fixed center and increasing width, similar to the eccentricity distribution of particles with nonzero initial eccentricity. However, Figure~\ref{F:inc}(b) shows different behavior: the distribution increasingly deviates from a normal distribution with time. The peak leans toward the mid-plane and asymmetry develops with more particles on the left side of the peak (i.e., smaller inclination). This indicates the diffusion is stronger near the mid-plane, which in fact is consistent with vertical stratification of the gas density. It would be interesting to measure the dependence of the diffusion coefficient on vertical height, but given our limited sample of initial inclinations, this remains to be investigated. We therefore focus our attention on the case of particles with zero initial inclination.
Figure~\ref{F:sigma_i} shows $\sigma_i(t)$ for disks with varying gravity parameter $\xi$ and particles with varying initial eccentricity $e_0$ but zero initial inclination $i_0 = 0$. Note that when $\sigma_i(t)$ is normalized by $\xi H / R$, all the curves roughly coincide. This indicates that $\sigma_i(t)$ is linearly dependent on both $\xi H / R$, but independent of $e_0$, similar to what we reported for the evolution of the mean radius and the eccentricity. The results can thus be summarized by the following expression:
\begin{equation} \label{E:sigma_i}
\sigma_i(t) = C_i\xi\left(\frac{H}{R}\right)\left(\frac{t}{P}\right)^{1/2},
\end{equation}
where $C_i$ is a dimensionless proportionality constant.
Table~\ref{T:orbit} lists our measured values of the constant $C_i$ for different resolutions and box dimensions when particles have zero initial inclination.
\placefigure{F:sigma_i}
We note that in contrast to the orbital radius and the eccentricity, the evolution of the orbital inclination is not significantly affected by the horizontal box size. The inclination of the particle orbits is mainly affected by the vertical structures of the density field, which in turn is related to the vertical scale height of the density stratification. The horizontal structures in the magneto-rotational turbulence may have little correlation with the vertical structures so that the horizontal box size has little effect on the vertical motions of the particles.
\subsection{Velocity Dispersion}
Finally, all three components of the velocity dispersion $\vec{\sigma}_u$ among the particles as a function of time can be measured in our stratified disk models. Similarly to what we reported in YMM09, each component assumes the same form:
\begin{equation} \label{E:sigma_u}
\sigma_{u,i}(t) = S_i\xi c_s \left(\frac{t}{P}\right)^{1/2},
\end{equation}
where the index $i$ is either $x$, $y$, or $z$ and $S_i$ is the corresponding dimensionless proportionality constant. We emphasize that $S_y \sim S_x / 2$ always holds because of the fixed ratio of amplitudes in $x$ and $y$ directions in the epicycle motions of the particles, which has been verified in our simulations. The values of $S_x$ and $S_z$ at different resolutions and box dimensions for the case of particles with zero initial inclination are listed in Table~\ref{T:orbit}. Note that the value of $S_x$ for the 2$\times$2$\times$4$H$ box at a resolution of 64~points/$H$ is consistent with that from the unstratified disk with the same horizontal size and resolution we reported previously (see Equation~(18) of YMM09), a further confirmation of the consistency between stratified and unstratified models discussed in Sections~\ref{SS:radius} and~\ref{SS:ecc}.
As shown by Figure~\ref{F:vel_disp}, we do not see evident dependence of velocity dispersion on resolution of our models. On the other hand, the horizontal component $S_x$ (and thus $S_y$) significantly depends on the horizontal box size $L_h$ while the vertical component $S_z$ does not. This is in accordance with the dependence of the three orbital properties found in previous subsections. From our measured values for $L_h \geq 4H$, we quantify $S_x$ and $S_z$ with the following expressions:
\begin{eqnarray}
S_x &\simeq& (1.0\times10^{-4})(L_h / H)^{1.08},\label{E:sx}\\
S_z &\simeq& 2.8\times10^{-4}.\label{E:sz}
\end{eqnarray}
\placefigure{F:vel_disp}
\subsection{Boundary Condition Effects} \label{SS:bce}
In order to confirm that our numerically convenient but perhaps physically inconsistent assumption of vertical field and periodic gas boundary conditions in the $z$-direction does not change our results significantly, we compare turbulence properties and particle dynamics from models with different vertical boundary conditions for the magnetic fields run at a lower resolution of 16 points per disk scale height $H$ in large boxes with size $16 \times 16 \times 4 H$. The gas density and velocity remain periodic. Figure~\ref{F:mri_bcz} shows that within about one disk scale height, the vertical structure of the rms density perturbation, the plasma $\beta$, and the $\alpha$ parameter are in quantitatively close agreement between the models. Note that the model with periodic boundary conditions creates a coronal region as low as $|z | \simeq 1.8H$, where $\beta \simeq 1$. More importantly, Figure~\ref{F:orbit_bcz} shows from top to bottom the time evolution in the standard deviations of particle orbital radius, eccentricity and inclination for the two models. The orbital radius and eccentricity agree very well irrespective of the vertical boundary conditions, and the effect on inclination differs by only $\sim$30\%, likely due to increased activity at high altitudes for the model with periodic magnetic fields. Thus, we believe that our results on particle dynamics are robust.
\placefigure{F:mri_bcz}
\placefigure{F:orbit_bcz}
\section{PROPERTIES OF THE STOCHASTIC FORCE EXERTED BY THE GAS} \label{S:force}
We have demonstrated in Section~\ref{S:turbulence} that various properties of the saturated magneto-rotational turbulence in our simulations converge with both resolution and box dimensions. However, while converging with resolution, the response of massless particles to the gravity of the turbulent gas does not similarly converge with the horizontal box size, as presented in Section~\ref{S:particles}. This result raises serious questions about the validity of using the local-shearing-box approximation to simulate the dynamics of any particles under gravitational influence of magneto-rotational turbulence. In order to understand the lack of convergence with box dimension in the orbital evolution of massless particles, we further study the properties of the gravitational force exerted by the turbulent gas.
\subsection{Magnitude and Fourier Amplitudes of the Force} \label{SS:force}
Table~\ref{T:force} lists the radial and the azimuthal components of the gravitational force exerted by the turbulent gas, $F_x$ and $F_y$, respectively. Since the force is stochastic, we ensemble average the root mean square of each force component over all particles, and then report it as time average and $1\sigma$ variation scaled by $\xi m_p H P$, where $m_p$ is the mass of a particle. Note that in our models, particles only follow the gravitational potential generated by the gas and the host star, so the acceleration of each particle is independent of its mass. This is an advantage in that the measurements reported in this paper can be used to gauge the gravitational influence of magneto-rotational turbulence on solid particles or planetary objects of any mass and thus its significance compared with other relevant interactions.
\placetable{T:force}
As can be seen in Table~\ref{T:force}, both the radial and the azimuthal components of the force generally increase with the horizontal box size. It is not clear if there exists any convergence in the radial component towards larger box size, since the 16$\times$16$\times$4$H$ box at a resolution of 32~points per $H$ is the only model that contradicts the general increasing trend. Certainly, the azimuthal component (i.e., the torque) shows no sign of convergence up to the largest box size we have investigated, with a regression power-law index of about 0.6. \cite{NG10} also reported in their recent local models a slight trend of increasing torque with increasing horizontal box size (see their Figure~8 and the corresponding discussion).
Since the density structure of the turbulent gas determines its gravitational force on the particles, it is informative to consider how the structure changes with resolution and horizontal box size. The density structure of the gas is best analyzed in Fourier space, using the Fourier amplitudes $\tilde{\rho}_1\left(k_x,k_y,z\right)$ defined in Section~\ref{SS:poisson}.\footnote{Note that $\tilde{\rho}_1\left(k_x,k_y,z\right)$ is strictly periodic in sheared coordinates and the wavenumber $\vec{k}$ is slightly different from that of a direct Fourier decomposition of $\rho_1(x,y,z)$; see Section~\ref{SS:poisson}.\label{F:wavenumber}} Figure~\ref{F:fourier} plots the time-averaged Fourier amplitudes of the gas density in the mid-plane $\tilde{\rho}_1\left(k_x,k_y,z=0\right)$ along either the radial or the azimuthal direction. The top-left panel shows the amplitude as a function of $k_x$ for $k_y = 0$. In general, the largest amplitude resides at the longest wavelength while monotonically decreasing with increasing wavenumber \citep{JYK09}. For any given box dimensions, increasing resolution has little effect on each amplitude, apart from extending the profile toward higher wavenumber. Flattening of the profile for wavelengths $\gtrsim$8$H$ is hinted for 16$\times$16$\times$4$H$ boxes. We find no evidence of a downward turn toward even longer wavelength. On the other hand, increasing the horizontal box size $L_h$ lowers the overall profile of the Fourier amplitudes. Interestingly, the amplitude of the longest wavelength mode $k_x = k_0 \equiv 2\pi / L_h$ remains roughly constant.
\placefigure{F:fourier}
The bottom-left panel of Figure~\ref{F:fourier} plots the summation of Fourier amplitudes over all $k_y$ at any given $k_x$. In contrast to the amplitudes for $k_y = 0$, the horizontal box size $L_h$ has little effect on the amplitudes for most of the wavelengths. Nevertheless, increasing resolution indeed increases the inertial range and resolves more power toward shorter wavelength.
The right-hand column of Figure~\ref{F:fourier} plots the azimuthal counterpart of the left-hand column. We find similar features in the azimuthal amplitudes as those found in the radial ones. The only difference is that the flat part of the spectrum in the azimuthal direction is short compared with that in the radial direction, as can be seen in the bottom-right panel. Power that is not captured in the long-wavelength range of the spectrum in the 2$\times$2$\times$4$H$ boxes is probably the cause for artificially higher amplitude at $(k_x,k_y) = (0,k_0)$ (the top-right panel), which may in turn be responsible for the anomaly found in the particle dynamics shown in Section~\ref{S:particles}.
The Fourier amplitudes shown in Figure~\ref{F:fourier} help us understand the relationship between the gravitational force exerted by the gas and the horizontal box size.
The radial and the azimuthal force components for any given (horizontal) Fourier mode $\vec{k} = (k_x, k_y)$ are related to the corresponding Fourier amplitude $\breve{\rho}(\vec{k})$ by (see Equation~\eqref{E:convol})
\begin{eqnarray}
F_x(\vec{k})
\sim 4\pi G\left(\frac{k_x}{k^2}\right)\left|\breve{\rho}(\vec{k})\right|
\lesssim 4\pi G\frac{\left|\breve{\rho}(k_0,0)\right|}{k_0}
\propto L_h,\label{E:fx}\\
F_y(\vec{k})
\sim 4\pi G\left(\frac{k_y}{k^2}\right)\left|\breve{\rho}(\vec{k})\right|
\lesssim 4\pi G\frac{\left|\breve{\rho}(0,k_0)\right|}{k_0}
\propto L_h,\label{E:fy}
\end{eqnarray}
where we have used the facts that the dominant modes in the radial and the azimuthal directions are the longest wavelength ones $\left|\breve{\rho}(k_0,0)\right|$ and $\left|\breve{\rho}(0,k_0)\right|$, and both are about constant irrespective of resolution and box size. Therefore, both components should be at most linearly proportional to the horizontal box size in our simulations. Although involving some simplifications, Equations~\eqref{E:fx} and~\eqref{E:fy} at least qualitatively explain the general trend of increasing stochastic force with horizontal box size.
\subsection{Correlation Time}
We next evaluate the correlation time $\tau_c$ of the stochastic torque. This quantity is in fact not well defined in the literature, and different authors have different preferences for its evaluation. For the purpose of consistency with the framework of radial diffusive migration of particles driven by random torque, we adopt the definition of \citet{JGM06}:
\begin{equation} \label{E:ct1}
\tau_c \equiv D(J) / \overline{\delta\Gamma^2},
\end{equation}
where $D(J)$ is the radial diffusion coefficient as a function of the Keplerian orbital angular momentum $J$, $\delta\Gamma$ is the fluctuating part of the torque, and the overline denotes an ensemble average. We repeat the relationship between $D(J)$ and the dimensionless parameter $\epsilon$ here \citep[see][and YMM09]{JGM06}:
\begin{equation} \label{E:diff_coeff}
D(J) = \frac{2.1\times10^{-3}}{16\pi}\epsilon\xi^2
\left(\frac{H}{R}\right)^2
\left(\frac{J^2}{P}\right).
\end{equation}
To calculate the correlation time $\tau_c$ with our definition, both the diffusion coefficient $D(J)$ and the magnitude of the stochastic torque are needed. As emphasized in YMM09, we have all the information about the particle distribution against each orbital property as a function of time, and thus we can directly evaluate the diffusion coefficient in the context of a diffusion equation, which is reported in Section~\ref{SS:radius} and Table~\ref{T:orbit}. Furthermore, we also have the ensemble-averaged rms value of the stochastic torque, which is reported in Section~\ref{SS:force} and Table~\ref{T:force}. Equations~\eqref{E:ct1} and~\eqref{E:diff_coeff} can then be used to calculate the correlation time $\tau_c$, and the result is listed in Table~\ref{T:force}.
As indicated in Table~\ref{T:force}, the correlation time of the stochastic torque $\tau_c$ does not appreciably depend on the resolution but it is significantly affected by the horizontal box size. The larger the horizontal box size, the longer the correlation time of the torque; this trend was also reported by \citet{NG10}. Up to the largest box size we have investigated ($L_h \lesssim 16H$), it is not clear if $\tau_c$ will converge with $L_h$. This behavior is similar to that of the torque magnitude, as discussed in Section~\ref{SS:force}.
As a side note, one may want to evaluate the correlation time of the stochastic torque $\tau_c$ without referring to stochastic particle orbital evolution and invert Equation~\eqref{E:ct1} to estimate the radial diffusion coefficient $D(J)$ indirectly. We have shown in YMM09, for the case of ideal unstratified disks, that the following formula, motivated by the definition of $D(J)$ under the framework of the Fokker-Planck formalism, serves this purpose:
\begin{equation} \label{E:ct2}
\tau_c \approx
\frac{\int_0^{\infty} \overline{\mathrm{ACF}(\tau)}\,\mathrm{d}\tau}
{2\overline{\mathrm{ACF}(0)}},
\end{equation}
where $\mathrm{ACF}(\tau)$ is the autocorrelation function of the stochastic torque as a function of the time lag $\tau$.
Here we repeat the exercise and use Equation~\eqref{E:ct2} to estimate $\tau_c$ for the case of ideal stratified disks. The result is listed in Table~\ref{T:force} and shows remarkable consistency with the values obtained directly from the evolution of particle orbital distribution. Therefore, Equation~\eqref{E:ct2} remains a useful estimator of the correlation time $\tau_c$.
To further understand the trend of increasing correlation time with increasing horizontal box size, we plot the autocorrelation function of the stochastic torque in Figure~\ref{F:acf}. As also demonstrated by \citet{NG10}, the larger the box, the less oscillatory the autocorrelation function is and the less pronounced is the negative contribution to the radial diffusion of the particles. This behavior in turn is consistent with increasing values of the correlation time and the radial diffusion coefficient.
\placefigure{F:acf}
In addition to the stochastic torque, a similar analysis may be conducted for the radial component of the gravitational force exerted by the gas. We find that the correlation time of the radial force also increases with the horizontal box size. However, our models indicate a much longer correlation timescale, which may amounts to several tens of orbital periods. Given that our simulations only lasted for about 100$P$, the magnitude of the correlation time is not well calculated and much longer simulations may be required to cover the full spectrum of the autocorrelation function for the radial force.
\citet{JYK09} have discussed the dominant role of the longest wavelength, purely radial Fourier mode $\vec{k} = (k_0,0)$ in great detail and suggested that these kind of persistent structures are similar to the zonal flows found in many other astrophysical systems. As noted by the authors, these axisymmetric structures can not generate torques that affect the orbital radius of the embedded particles. While the radial forces due to the zonal flows can still affect the eccentricity of the particles, their long correlation time indicates that they may not be responsible for the diffusive evolution of the particle eccentricity, as shown in YMM09 and Section~\ref{SS:ecc}.\footnote{It remains possible that the stochastic radial forces due to small-scale axisymmetric density perturbations drive the diffusive evolution of the particle eccentricities.} On the other hand, the longest wavelength, purely azimuthal Fourier mode $\vec{k} = (0, k_0)$ in the gas structure\footnote{Note that this mode does not represent a single shearing wave, and the wavenumber $(k_x,k_y)$ is independent of time; see footnote~\ref{F:wavenumber} and Section~\ref{SS:poisson}.} is significant enough that both the radial diffusive migration and the eccentricity evolution of the particles are affected by the large-scale azimuthal density perturbations and thus the box dimensions.
\section{RELATION OF BOX SIZE TO PARTICLE ORBITAL EVOLUTION} \label{S:box_size}
As described in Section~\ref{S:force}, the magnitude and the correlation time of the gravitational force exerted by the turbulent gas correlates with the horizontal size of our local shearing box. This behavior essentially explains the correlation of the stochastic evolution of the particle orbital radius and eccentricity with the horizontal box size reported in Section~\ref{S:particles}. These results challenge the validity of the local shearing box to describe large-scale density structures in magneto-rotational turbulence.
By measuring the two-point correlation functions of density, velocity, and magnetic field in magneto-rotational turbulence, \citet{GG09} found extended density features in contrast to well localized magnetic structures. The authors suggested that the density features are propagating acoustic waves excited by the turbulence \citep[see also][]{HP09a,HP09b}. It is not clear yet what the dissipation length scale for these waves is, which is crucial to understanding whether a local shearing box can capture the largest-scale structures in the density fluctuations excited by magneto-rotational turbulence.
On the other hand, a global disk model may require high resolution in order to self-consistently produce large-scale structures. For example, \citet{JYK09} argued that the inverse cascade of magnetic energy from small scales to large scales might be responsible for ultimately launching zonal flows. To confirm this mechanism in a global context, a model that is capable of resolving at least the correlation lengths in the magnetic structures might be necessary.
\subsection{Bridge to a Global Disk Model?}
Given the correlation with horizontal box size reported in this work, it is not clear how a local shearing box can be connected to a global disk model so that both models give consistent results on the particle dynamics under the gravitational influence of density fluctuations in magneto-rotational turbulence.
Nevertheless, we conjecture at this point that the criterion $L_h / R \sim O(1)$ might provide a physical length scale for a local shearing box. At this scale, the local-shearing-box approximation formally breaks down since it assumes $L / R \ll 1$. The curvature terms become important and this might trigger turbulent eddies to damp propagating waves. If this conjecture could be verified, Equations~\eqref{E:cx}, \eqref{E:ce} and~\eqref{E:sx} would prove useful in evaluating the gravitational influence of the turbulent gas on particle dynamics. Substituting $R$ for $L_h$ in these equations, our models predict that
\begin{equation} \label{E:cs_predict}
\left\{
\begin{array}{ccl}
C_x &\simeq& 6.6\times10^{-5}\,(H/R)^{-1.35}\\
C_e &\simeq& 7.2\times10^{-5}\,(H/R)^{-1.08}\\
S_x &\simeq& 1.0\times10^{-4}\,(H/R)^{-1.08}
\end{array}
\right.\qquad \textrm{for $\alpha \sim 10^{-2}$.}
\end{equation}
\citet{NG10} presented the most recent work on stochastic planetesimal orbital evolution in a global disk model. Their measurement of the orbital radius and radial velocity dispersion can be directly compared with ours. These authors reported in their global model G3, whose $\alpha\simeq0.02$ was the closest to ours, that $\sigma_x(t) = 7.2\times10^{-4}\,R\sqrt{t/P}$ and $\sigma_{u,x}(t) = 8.2\times10^{-3}\,c_s\sqrt{t/P}$. In this specific model, a disk aspect ratio\footnote{See footnote~\ref{scale_height}.} of $H / R \approx 0.071$ and a strength of gas gravity of $\xi \approx 2.1$ at 5~AU were adopted. In comparison, Equations~\eqref{E:sigma_x}, \eqref{E:sigma_u}, and~\eqref{E:cs_predict} give $\sigma_x(t) \approx 6.9\times10^{-3}\,(H/\sqrt{2})\sqrt{t/P} = 3.4\times10^{-4}\,R\sqrt{t/P}$ and $\sigma_{u,x}(t) \approx 3.8\times10^{-3}\,c_s\sqrt{t/P}$. Our results are only about a factor of two lower than those of \citet{NG10}. Therefore, the criterion $L_h \sim R$ for a local shearing box seems to offer some consistency between global and local models.
These comparisons are not meant to be accurate and conclusive, though, since the proposed criterion $L_h / R \sim O(1)$ invites significant uncertainty. Further investigation into the large-scale structures of magneto-rotational turbulence and the discrepancy between global and local disk models is still needed.
\subsection{Implications for Planet Formation and Migration}
If the criterion $L_h \sim R$ holds, Equations~\eqref{E:epsilon} and~\eqref{E:gamma} imply
\begin{equation} \label{E:eg_predict}
\left\{
\begin{array}{ccl}
\epsilon &\simeq& 6.5\times10^{-6}\,(H / R)^{-2.69}\\
\gamma &\simeq& 3.6\times10^{-5}\,(H / R)^{-1.08}
\end{array}
\right.\qquad\textrm{for}\ \alpha \sim 10^{-2}.
\end{equation}
Consider a region at $R = 5$~AU in a minimum-mass solar nebula \citep[MMSN;][]{cH81}, where $H / R \approx 0.08$. Equations~\eqref{E:eg_predict} give $\epsilon \approx 7\times10^{-3}$ and $\gamma \approx 6\times10^{-4}$. This value of the $\epsilon$ parameter is about one order of magnitude higher than was reported in YMM09 for the case of a 2$\times$2$\times$2H unstratified box, while the value of the $\gamma$ parameter is only about a factor of three larger. Note that according to Equations~\eqref{E:eg_predict}, both $\epsilon$ and $\gamma$ decrease with increasing disk aspect ratio $H / R$, which often increases with increasing radial distance $R$.
We suggested in YMM09 that in a typical protoplanetary disk, the radial diffusive migration of protoplanets induced by magneto-rotational turbulence may be unimportant compared to secular migration. According to \citet{JGM06}, for the diffusive migration to be able to dominate over type I migration, the $\epsilon$ parameter should be greater than or about 0.1--1 for the MMSN. Our new measurement remains orders of magnitude smaller than this transition value. Therefore, our previous conclusions on the insignificance of planetary diffusive migration may still hold even though $\epsilon$ is increased by an order of magnitude. However, note that recent calculations for type I migration in non-isothermal disks \citep[e.g.,][]{LPM10,PB10,PB11} may negate the dominance of secular migration over stochastic migration driven by magneto-rotational turbulence.
We also argued in YMM09 that kilometer-sized planetesimals moving in magneto-rotational turbulence survive mutual collisional destruction, except in the inner region of a young protoplanetary disk. This was based on the results of \citet{IGM08} for the cases of $\gamma = 10^{-3}$ and $10^{-4}$. Since our new measurement does not fall outside this range, the same conclusion still applies.
In contrast, \citet{NG10} argued that the velocity dispersion of kilometer-sized planetesimals excited by magneto-rotational turbulence might be so large that these planetesimals should be erosive to each other. Before a consistent scenario for the survivability of planetesimals could be assembled, however, several factors remain to be accounted for. First, eccentricity damping due to tidal interaction between a planetesimal and its surrounding gas acts to circularize the orbits of the planetesimals \citep[e.g.,][]{GT80}. This mechanism was absent in the models of \citet{NG10} and thus their measurement of the velocity dispersion of planetesimals should be considered as an upper limit.\footnote{This is not true if mutual gravitational scattering between planetesimals dominates over turbulence-driven excitation. However, this might not be the case in a typical protoplanetary environment; see YMM09.} On the other hand, using analytical arguments to estimate the equilibrated eccentricity of planetesimals, \citet{IGM08} took three possible eccentricity damping effects into account --- tidal interaction, gas drag, and inelastic collisions --- and thus their estimate of the velocity dispersion of planetesimals might be more realistic. Even though significant uncertainty was involved in the assumed strength of turbulent excitation of eccentricity in the models of \citet{IGM08}, it could be improved by more accurate measurement of the $\gamma$ parameter, as provided by the present work.
Other interesting progress in the study of mutual collisions of planetary objects includes different outcomes from different impact angles \citep[head-on vs.\ hit-and-run scenarios; see, e.g.,][]{eA09,MS09} and improved calculations on material properties \citep[e.g.,][]{LS09,SL09,jW10}. Given all these new developments, a reconsideration of the survivability of kilometer-sized planetesimals moving in magneto-rotational turbulence seems warranted.
\section{SUMMARY} \label{S:summary}
Directly extending our previous publication (YMM09), we continue to study massless particles moving under the gravitational influence of density fluctuations due to saturated magneto-rotational turbulence in a local, isothermal, Keplerian gas disk. We include linearized vertical gravity from the host star and thus vertical stratification of the gas disk. For comparison, the conditions in the mid-plane of the vertically stratified disks are exactly the same as those in the unstratified disks of YMM09.
In order to accurately measure the gravitational effect of the turbulent gas, we separate the gas density $\rho$ into two components: the basic state $\rho_0(z)$ for the vertical hydrostatic equilibrium (Equation~\eqref{E:rho_0}) and the density deviation $\rho_1 \equiv \rho - \rho_0$ from this basic state. We use the exact gravitational acceleration due to the basic state (Equation~\eqref{E:g_0}) and only solve the Poisson equation for the gravitational potential due to the density deviation (Equation~\eqref{E:poisson}). We emphasize that since the Poisson equation is linear in density, this approach does not assume small density fluctuations. Furthermore, we implement isolated boundary conditions in the vertical direction and thus any density fluctuation outside the vertical computational domain is neglected.
By imposing a weak, uniform external magnetic field, we maintain a constant level of saturated magneto-rotational turbulence in the disk mid-plane. Several turbulence properties demonstrate convergence with both resolution up to 64~points per disk scale height $H$ and horizontal box size up to 16$H$. The \citet{SS73} $\alpha$ parameter in the mid-plane of our models is controlled at the level of $\sim$10$^{-2}$.
However, even though the properties of the turbulent gas appear numerically convergent, the dynamics of massless particles moving under the gravity of this turbulent gas does not converge with the horizontal box size $L_h$. The larger the horizontal box size, the stronger the gravitational effect of the gas on the particles. Specifically, the evolution of the orbital radius, the eccentricity, and the horizontal velocity dispersion of the particles is roughly linearly dependent on $L_h$ up to 16$H$. This trend was also found in our unstratified models (YMM09), and we find consistency between the unstratified models and the mid-plane of the stratified models. In contrast to the horizontal components of the particle movement, we find that the evolution of the inclination and the vertical velocity dispersion is not significantly affected by $L_h$.
The dependence of particle dynamics on the horizontal box size can be traced back to the density structure of the gas. Consistent with the large-box models studied by \citet{JYK09}, the longest wavelength Fourier mode dominates the density spectrum along the radial direction in our models. Furthermore, we find that the longest wavelength Fourier mode in the azimuthal direction also strongly influences the particle dynamics, leading to the diffusive evolution of both the orbital radius and the eccentricity of the particles. The spectral amplitudes of these longest wavelength modes are roughly constant against the horizontal box size. Using a simple single-mode analysis, we show that the linear dependence of the particle response is a natural outcome of these findings for the density spectrum.
Correlation of particle dynamics with box size poses a major difficulty for the interpretation of local-shearing-box models involving gravitational physics of magneto-rotational turbulence. We can nevertheless conjecture that $L_h \sim R$, where $R$ is the distance of the box center to the host star, might be a natural scale of choice for a local model to approach reality. If this conjecture holds, we find that our previous conclusions in YMM09 on the unimportance of radial diffusive migration for protoplanets as well as the survivability of kilometer-sized planetesimals under collisional destruction may still be valid. Ultimately, high-resolution global disk models and detailed comparisons with large-box local models might be necessary to settle this issue.
\acknowledgments
We thank the anonymous referee for critical comments that significantly improved the rigor and clarity of this paper, and thank Charles Gammie, Anders Johansen, Roman Rafikov, Cl\'{e}ment Baruteau, Martin Pessah, Richard Nelson, and Oliver Gressel for their insightful discussion on this research. This research was supported in part by the Perimeter Institute for Theoretical Physics. Resources supporting this work were provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center. Partial support of this work was provided by the NASA Origins of Solar Systems Program under grant NNX07AI74G.
|
\section{The $p$-median problem}
The challenge in facility location problems is to place $p$ service centers
or facilities so that $n$ demand points are optimally served (see for example
Ref.~\onlinecite{DreznerHamacher02} for an overview).
Facilities can be hospitals, supermarkets, fire stations, libraries,
warehouses, or any other supply centers providing vital resources to the
population living at the demand points (e.g., households or cities).
Here we consider the case where the demand points are at regular intervals
along a one-dimensional geographic object, such as a road or a river, and
where every demand point is a possible location for a facility.
The number of people $N$ who require the facilities' services is assumed to
be known at each demand point.
This number is typically very heterogeneous across geographic space.
Depending on the context, there are different strategies for the placement
of the facilities.
In this article, we concentrate on the $p$-median problem, an important
special case, where the objective is to minimize the average distance
between a person's demand point and the nearest facility.
(A recent summary of the vast literature on the $p$-median problem can be
found in Ref.~\onlinecite{Reese06}).
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{pmed_intro.eps}
\caption{Illustrative example of the $p$-median problem along a line.
The population $N$ is known at the demand points $q_1, \ldots, q_n$.
In this article, the distance between neighboring demand points is
assumed to be constant.
Facilities will be placed on $p$ of these $n$ demand points.
(In the figure, $p=4$.)
Their locations $r_1, \ldots, r_p$ are to be determined so that the
average distance between a demand point and the nearest facility,
weighted by $N$, is minimized.
After the facilities have been located, the line can be divided into $p$
segments $s_1, \ldots, s_p$ so that the $i$-th segment corresponds to the
service region of the $i$-th facility.
}
\label{pmed_intro}
\end{center}
\end{figure}
Let us call the facility locations from left to right $r_1, \ldots, r_p$.
These positions are chosen among the demand points $q_1, \ldots, q_n$, which
are equidistant (i.e., $q_{i+1}-q_i = \text{const.}$ for $i=1, \ldots, n-1$)
along a line (see Fig.~\ref{pmed_intro}).
If the population at $q_i$ is denoted by $N_i$, the $p$-median problem consists
of minimizing the cost function~\footnote{
Because $N_i$ for a given location problem is constant, we could in principle
directly minimize the numerator in Eq.~\ref{pmedian} and ignore the
denominator. We have decided to keep the denominator so that $C$ equals the
average distance. $C$ can then be more easily compared across
different location problems.
}
\begin{equation}
C(r_1,\ldots,r_p) =
\frac{\sum_{i=1}^{n} N_i \min_{j=1,\ldots,p}|q_i-r_j|}{\sum_{i=1}^{n} N_i}.
\label{pmedian}
\end{equation}
Because only trips to the nearest facility play a role in Eq.~\ref{pmedian},
the line along which the demand points are located can be partitioned into
$p$ segments or service regions.
Demand points belong to the same segment if and only if they share the same
closest facility, see Fig.~\ref{pmed_intro}.
The length of facility $i$'s service region is given by
\begin{equation}
s_i =
\begin{cases}
\frac12 (r_1 + r_2)-q_1 & \text{if $i=1$},\\
\frac12 (r_{i+1} - r_i) & \text{if $i=2, \ldots, p-1$},\\
\frac12 (r_{p-1} + r_p)-q_n & \text{if $i=p$}.
\end{cases}
\end{equation}
We will now take a closer look at the relation between $s_i$ and the
population density around facility $i$.
\section{Scaling of the lengths of the service regions}
At first sight, it is plausible that the spatial density of facilities should
follow the same trend as the population density: where there are more people
there should be proportionately more facilities.
However, as we will see shortly, the $p$-median solution does not follow this
rule that would give every facility an equal number of customers.
Instead facilities are less abundant per capita in the high-demand regions
than in the low-demand regions.
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{cont_approx.eps}
\caption{(a) Under the assumption that the population $N$ (gray histogram)
varies little between neighboring demand points, $N$ can be approximated
by a continuous function $\rho$ (black curve).
(b) The function $\sigma(x)$ is defined as the length $s$ of the
segment covering position $x$.
Strictly speaking, $\sigma$ is a piecewise constant function.
However, if the spatial variations in $N$ are sufficiently small,
$\sigma$ can be approximated by a continuous function
(indicated by the dotted curve).
}
\label{cont_approx}
\end{center}
\end{figure}
For a spatially heterogeneous population distribution $N_i$, it is difficult
to deduce this general trend directly from Eq.~\ref{pmedian}.
With certain approximations, however, the problem becomes analytically
tractable; essentially, we translate the line of reasoning developed in
Ref.~\onlinecite{Gusein93} and \onlinecite{GastnerNewman06} for the
two-dimensional $p$-median problem to the one-dimensional case.
First we define the population density $\rho(x)$ which is the number of people
per unit length in the vicinity of $x$.
Equation~\ref{pmedian} can be rewritten as
\begin{equation}
C(r_1, \ldots, r_p) =
\frac{\int_{q_1}^{q_n} \rho(x)\min_{j=1,\ldots,p}|x-r_j|\,dx}
{\int_{q_1}^{q_n} \rho(x)\,dx},
\label{pmed_cont_approx}
\end{equation}
where we have used the new notation to replace sums by integrals.
If we allow $\rho$ to be piecewise constant, this expression is still
exact, but later it will be more convenient to approximate
$\rho$ with a continuous function (Fig.~\ref{cont_approx}a).
Next we define $\sigma(x)$ to be the length of the
segment serviced by the facility closest to $x$ (see Fig.~\ref{cont_approx}b).
The average distance from facility $j$ to a point $x$ inside its service
region is equal to $g_j \sigma(x)$, where $g_j$ depends on the exact location of
the facility.
For example, if $r_j$ is close to the center of the segment,
$g_j\approx\frac14$.
In the spirit of a mean-field approximation, we will now assume that $\rho$
varies little over the size of a segment.
Then we can replace the exact distance, $\min|x-r_j|$, in the numerator of
Eq.~\ref{pmed_cont_approx} with its average $g_j\sigma(x)$,
\begin{equation}
C \approx \frac{\int_{q_1}^{q_n} \rho(x)\,g\,\sigma(x)\,dx}
{\int_{q_1}^{q_n} \rho(x)\,dx}.
\label{pmed_mf_approx}
\end{equation}
The index $j$ was dropped in Eq.~\ref{pmed_mf_approx} assuming that most
facilities will be close to the center of their service region so that
$g_j$ is approximately constant.
Unlike in Eq.~\ref{pmedian}, the locations $r_j$ no longer appear explicitly
in Eq.~\ref{pmed_mf_approx}.
Instead we have to find the function $\sigma(x)$ that minimizes $C$ subject to
the constraint that there are $p$ facilities.
This constraint can be expressed as
\begin{equation}
\int_{q_1}^{q_n} \frac1{\sigma(x)}dx = p.
\label{constraint}
\end{equation}
Introducing a Lagrange multiplier $\alpha$, the problem is equivalent to
finding the zero of the functional derivative
\begin{equation}
\frac{\delta}{\delta \sigma}
\left[\frac{g\int_{q_1}^{q_n} \rho(x)\,\sigma(x)\,dx}
{\int_{q_1}^{q_n} \rho(x)\,dx}
-\alpha\left(p-\int_{q_1}^{q_n}\frac1{\sigma(x)}dx\right)\right] = 0,
\end{equation}
solved by
\begin{equation}
\sigma(x) = \sqrt\frac{\alpha\int_{q_1}^{q_n}\rho(x')\,dx'}{g\,\rho(x)}.
\end{equation}
The Lagrange multiplier can be eliminated by inserting this expression into
Eq.~\ref{constraint}.
After some algebra,
\begin{equation}
\sigma(x) = \frac{\int_{q_1}^{q_n} \sqrt{\rho(x')} dx'}{p\sqrt{\rho(x)}}
\propto [\rho(x)]^{-1/2}.
\label{sqrt_scaling}
\end{equation}
The lengths of the service regions are thus inversely proportional to the
square root of the population density.
The spatial density of facilities $1/\sigma$ increases $\propto \rho^{1/2}$,
but the per-capita density $1/(\rho\,\sigma)$ decreases $\propto \rho^{-1/2}$
with growing population.
The square-root scaling is a compromise providing most services where they
are most needed, namely in the densely populated regions, but still leaving
sufficient resources in sparsely populated regions where travel distances are
longer.
This result implies an economy of scales: In crowded cities
fewer facilities per capita can supply a larger population than in rural
areas.
If facilities and demand points are not restricted to be along a line, but
can be placed in two-dimensional space, the scaling exponent is $2/3$
instead
of $1/2$~\cite{Gusein93,GastnerNewman06} (see Section~\ref{conclusion}).
However, economies of scale are also predicted in two dimensions.
Empirical studies have indeed reported this effect for certain classes of real
facilities~\cite{Stephan77,Bettencourt_etal07,Um_etal09}.
\section{Exact solution for empirical population distributions}
\label{exact_numerical}
The calculation in the previous section assumes that the population density
$\rho(x)$ varies little within a service region.
As we can see from Eq.~\ref{sqrt_scaling}, this implies that the segment
length $\sigma(x)$ is also a smooth function (Fig.~\ref{cont_approx}b).
Real census data, however, typically reveal strongly varying populations
even on small spatial scales.
In Fig.~\ref{test_sets}a--d, we show population numbers near three US
Interstate highways and the navigable Mississippi River.
The data were generated from the US census of the year 2000.
First, Interstates 5, 10, 90 and the Mississippi River were
parameterized by arc length and markers were placed at regular 1-km intervals.
Then census blocks within 10 km of the highways or the Mississippi were
identified and their population assigned to the nearest kilometer marker.
As is clear from Fig.~\ref{test_sets}a--d, neither of the four populations is
a smooth function.
Whether the assumptions behind Eq.~\ref{sqrt_scaling} are valid, is questionable,
but it turns out that the scaling law for the service regions still holds with
surprising accuracy.
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{test_sets.eps}
\caption{Population $N$ as a function of position $x$ along (a)
Interstate 5, (b) 10, (c) 90, (d) the navigable part of the
Mississippi River.
The small squares below the $x$-axes indicate the optimal $p$-median
positions of $100$ facilities.
(e) Map of the roads, the river, and the facility locations.
}
\label{test_sets}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{scaling_exact_opt.eps}
\caption{(Color online) The length of a service region $s$ versus
the mean population $\langle N\rangle$ in this service region.
Lines indicate least-squares fits to Eq.~\ref{log-log-scaling}.
Scaling is in good agreement with the analytic prediction
$s\propto\langle N\rangle^{-1/2}$.}
\label{scaling_exact_opt}
\end{center}
\end{figure}
To compute the scaling exponent, $p=100$ facilities are placed on each of the
four test data sets.
The optimal locations are calculated with the efficient algorithm of
Ref.~\onlinecite{HassinTamir91}.
Their positions along the roads and the river in geographic space are shown
in Fig.~\ref{test_sets}e.
The segment lengths $s_i$ are calculated for each facility $i=0,\ldots,p$.
In Fig.~\ref{scaling_exact_opt}, $s_i$ is plotted versus the mean value of
$N$ inside the segment, denoted by $\langle N\rangle_i$
\footnote{If both the left and the right boundary, $b_l$ and $b_r$, of the
segment are half-integers, $\langle N\rangle_i$ is defined as
$\sum_{j=\lceil b_l+1/2\rceil}^{\lfloor b_r-1/2\rfloor}N_j/s_i$
If $b_l$ or $b_r$ is an integer, $N_{b_l}/(2s_i)$ or $N_{b_r}/(2s_i)$ is
added to the sum (i.e., half of the population is assigned to the facility on
the right, half to the equally distant facility on the left).
}.
Ordinary least-squares fits of
\begin{equation}
\log(s_i) = a\log\langle N\rangle_i + \text{const.}
\label{log-log-scaling}
\end{equation}
to the data yield slopes $a=-0.511$ (I-5), $-0.514$ (I-10), $-0.504$ (I-90),
and $-0.496$ (Mississippi River), close to the prediction $a=-1/2$ of
Eq.~\ref{sqrt_scaling}.
The correlations are strong; $R^2$ is consistently bigger than $0.89$.
Assuming that the residuals are log-normally distributed, the predicted
value $-1/2$ is in all cases within the 95\% confidence intervals.
Thus, the equivalent of Eq.~\ref{sqrt_scaling},
$s_i \propto (\langle N\rangle_i)^{-1/2}$,
obtained by replacing the continuous variables $\rho$ and $\sigma$ by their
discrete counterparts $\langle N\rangle_i$ and $s_i$, is a good approximation.
This observation demonstrates that scaling at the exact $p$-median
configuration is robust
even in the presence of strong spatial fluctuations.
\section{The number of configurations for non-minimal costs}
That the square-root scaling of the service regions is discernible even
for realistically heterogeneous input, establishes a potential link to
previous empirical work.
Data collected in Ref.~\onlinecite{Stephan88,Bettencourt_etal07,Um_etal09}
suggest, at least for certain classes of facilities, a sublinear dependence
of service facilities on population numbers.
It has been conjectured that the $p$-median model~\cite{Stephan77} or a
generalization thereof~\cite{Stephan88,Um_etal09} might explain this trend.
Admittedly, we are looking in this article at a simplified linear geometry.
Yet that sublinear scaling is robust even for substantially noisy input,
might be viewed as supporting evidence for this conjecture.
However, there is more to the problem than first meets the eye.
Although it is mathematically convenient to assume that facilities are
placed to minimize an objective function such as Eq.~\ref{pmedian}, it is far
from clear that the exact minimum will be achieved in reality.
Decisions about facility locations are probably more haphazard in real life.
For example, site selections may be swayed by political interests, short-term
fluctuations in property prices, or based on an incomplete knowledge of
the actual demand.
Even if the best effort is made to reach the global optimum, ``accidents of
history'' may keep the facility locations trapped in a costlier local optimum.
It seems overly optimistic to draw conclusions about the scaling of real
service regions only from the best of all solutions.
The available literature for real facility
distributions~\cite{Stephan88,Bettencourt_etal07,Um_etal09} -- rather than
the numerically optimal ones discussed in Sec.~\ref{exact_numerical} --
also justifies cautious skepticism, as some significant differences to the
$p$-median result have been observed in reality, albeit in two dimensions.
How many facility configurations with costs near, but not necessarily equal to,
the global minimum exist?
There is no simple way to answer this question.
Although the algorithm of Ref.~\onlinecite{HassinTamir91} can find the global
optimum very efficiently, it does not provide information about non-optimal
solutions.
Scanning all possible configurations is out of the question because their
number is too vast.
Even for our smallest test data set (I-5) there are
$\binom{2213}{100}\approx 3.5\cdot 10^{175}$ different ways to locate the
facilities.
The situation is reminiscent of many-particle systems in physics where one
wishes to calculate the large number of micro-states at a certain energy
level out of an even larger number of all conceivable micro-states.
In that context, statistical mechanics has developed many powerful numerical
tools.
We will build on this analogy in order to estimate the number of
non-optimal facility locations.
Let us call $\Omega(C)dC$ the number of facility locations with costs between
$C$ and $C+dC$.
The function $\Omega(C)$ plays the role of the ``density of states'' in
statistical mechanics.
As we will see, $\Omega$ increases very rapidly as $C$ exceeds the minimum
$C_{min}$, so that it will be more convenient to work with its logarithm,
the entropy $S(C)=\log\Omega(C)$.
Our aim is to calculate $S$ with Monte Carlo simulations.
Several methods exist~\cite{Lee93,deOliveira_etal96,Argollo_Lima03};
here we apply the Wang-Landau algorithm~\cite{WangLandau01a}.
First, the range of possible costs is divided into small discrete intervals of
length $\Delta C$.
Then a random walk through the set of facility locations is performed and
we count, in the form of a histogram, how often each interval is visited.
The main idea behind the Wang-Landau algorithm is to bias the random walk in
such a manner that all intervals are visited equally often.
For such a ``flat histogram'' we obtain equally good statistics for all
intervals, an advantage when $S(C)$ is the basis of further calculations.
We describe details of our implementation in App.~\ref{wang_landau}.
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{entropy.eps}
\caption{(Color online) The entropy $S$ (i.e., the logarithm of the density of states)
versus the cost $C$.
The inset shows the same four curves as the main panel, but with
rescaled abscissa $(C-C_{min})/n$ where $n$ is the number of
demand points.}
\label{entropy}
\end{center}
\end{figure}
From calculations for four different empirical population distributions
(Fig.~\ref{entropy}) it is clear that $S$ is singular at $C_{min}$, the
smallest possible cost.
Thus, $S$ increases enormously in the vicinity of $C_{min}$ and the density
of states $\Omega=\exp(S)$ grows even more rapidly.
The results for four different empirical population distributions suggest that
$S$ follows approximately the same curve (inset of Fig.~\ref{entropy}) if
regarded as a function of $(C-C_{min})/n$, where $n$ is the total number of
demand points $q_1,\ldots,q_n$.
Therefore, it appears to be a universal feature that for all realistic
populations a large number of different possible configurations must be
considered if the assumption of optimality is relaxed.
This observation raises the question: Can the scaling relation of
Eq.~\ref{sqrt_scaling} still be observed if facility locations are not
exactly optimal, but are among the numerous configurations achieving
almost but not exactly $C_{min}$?
\section{Is scaling detectable for non-minimal costs?}
If we randomly select a facility configuration with a cost in the interval
$[C,C+dC]$, we can formally obtain the scaling exponent $a$ from
Eq.~\ref{sqrt_scaling} as follows.
First, we log-transform the segment lengths $s_i$ and the population density
$\langle N\rangle_i$.
Then a least-squares linear fit to Eq.~\ref{log-log-scaling} will be performed
to calculate $a$.
This procedure can be coupled with the Wang-Landau algorithm so that, at
every step in the random walk through configuration space, we compute $a$,
the cost $C$, and at the end of the algorithm the mean value
$\langle a\rangle$ as a function of $C$.
\begin{figure}
\begin{center}
\includegraphics[width=8.6cm]{exp_rsq.eps}
\caption{(Color online) (a) The mean scaling exponent $\langle a\rangle$, (b) the
coefficient of determination $R^2$ as a function of the cost $C$.}
\label{exp_rsq}
\end{center}
\end{figure}
The results, shown in Fig.~\ref{exp_rsq}a, indicate that $\langle a\rangle$ is
approximately $-1/2$ at the minimum cost $C_{min}$ for all four numerical
test sets, as anticipated by our earlier calculations.
As the cost increases, $a$ also increases, indicating a decreasing dependence
of the segment lengths on the population.
This behavior makes sense because the facility locations become more random
as we move away from the optimum.
Interestingly, the overall trend how $\langle a\rangle$ increases with $C$
is similar in all four cases.
In particular, the behavior near the minimum is noteworthy because
$\langle a\rangle$ increases most rapidly near $C_{min}$.
In other words, the analytic prediction at $C_{min}$ -- which provides us
with the only easily calculable reference point for $a$ -- is unfortunately
at the point where small deviations can also cause the greatest changes in
$a$.
Together with the least-squares exponent $a$, we can also obtain other
statistical measures from linear regression, such as the coefficient of
determination $R^2$ (Fig.~\ref{exp_rsq}b).
It can take values between $0$ and $1$; the higher its value, the stronger
the correlation between $s$ and $\langle N\rangle$.
In our numerical test sets, $R^2$ takes its highest value ($\approx 0.9$) at
$C_{min}$ and decreases as we move toward higher costs following a slightly
sigmoidal curve toward values around zero.
At very large costs, $R^2$ increases again because the solution is
effectively an ``obnoxious facility'' location where facilities are in
sparsely populated regions so that $s$ and $\langle N\rangle$ are positively
(instead of negatively) correlated.
For costs near $C_{min}$, however, an increase in $\langle a\rangle$ is
coupled with a reduction in $R^2$.
\section{Conclusion}
\label{conclusion}
In this article, we have studied the one-dimensional $p$-median problem.
In one dimension, the exact optimum can be calculated numerically in
polynomial time; in two dimensions~\cite{MegiddoSupowit84} and on arbitrary
graphs~\cite{KarivHakimi79}, the $p$-median problem is NP-complete so that no
polynomial-time algorithm is currently known.
Previous empirical studies of scaling in real facility locations have usually
dealt with two-dimensional densities.
The approximate analytic result in one dimension, $\sigma\propto\rho^{-1/2}$
(Eq.~\ref{sqrt_scaling}), can be easily generalized for arbitrary
dimension $d$, where the size of a $d$-dimensional Voronoi cell $\sigma$ is
predicted to scale as $\rho^{-d/(d+1)}$.
The scaling of the facility density $1/\sigma$ with the population density
$\rho$ thus remains sublinear in all dimensions.
Numerical optimization in two dimensions, based on US census data, yields
indeed an exponent in excellent agreement with the predicted exponent
$a=-2/3$~\cite{GastnerNewman06}.
In 1977, Stephan implicitly proposed that the $p$-median model might explain
empirical scaling relations between the area and population density of
subnational administrative units (e.g., states, provinces,
counties)~\cite{Stephan77}.
Although he later generalized the objective function as more data became
available~\cite{Stephan88}, the notion that facilities may self-organize
towards sublinear scaling has remained attractive, as proved by the recent
rediscovery of Stephan's model by Um et~al.~\cite{Um_etal09}.
However, as the work shown here underlines, one has to be careful when
interpreting empirical data.
Increased spatial noise in the facility distribution can lead to different
exponents and reduced correlations.
The situation investigated here portrays only one special scenario how
randomness might be present, namely as a uniform probability distribution
over all costs in an interval $[C,C+dC]$.
It is also conceivable that not all configurations within this range are
equally likely, so that the best-fit exponents may behave differently.
We may also replace the $p$-median model by a different optimization principle
(e.g., competitive facility location such as the Hotelling
model~\cite{Hotelling29}) which can change the exponent at the optimum.
However, we believe that a steep increase in the number of possible
configurations is a generic tendency of most models that relax the constraint
of strict optimization even to a small degree.
\section{Acknowledgments}
The author thanks M.~E. Moses, S. Banerjee, B. Blasius, and H. Youn for
stimulating discussions. The author acknowledges support from Imperial College.
|
\section{Introduction}
\label{sec:intro}
Reducing scanning time in Magnetic Resonance Imaging~(MRI) exams remains a worldwide challenging issue.
The expected benefits of a faster acquisition can be summarized as follows : i) limit patient's exposure to the MRI environment either for safety or discomfort reasons,
ii) maintain a strong robustness in the acquisition with respect to subject's motion artifacts,
iii) limit geometric distortions or maintain high image quality and iv) acquire more spatially or temporally resolved images in the same or even reduced amount
of time~\cite{Kochunov05b,Rabrait07}.
The basic idea to make MRI acquisitions faster or to improve spatial resolution at fixed scanning time
consists of reducing the amount of acquired samples in the $k$-space and developing dedicated reconstruction pipelines.
To achieve this goal, two main research avenues have been developed so far:
i) {\em parallel imaging} that relies on a geometrical complementarity principle involving multiple receiver channel coils with complementary sensitivity profiles.
This enables the reduction of the number of $k$-space lines to be acquired without degrading spatial resolution or truncating the Field of View~(FOV), and
thus requires the combination of reduced FOV coil-specific images to reconstruct the full FOV image~\cite{pruessmann_99,griswold_02}.
ii) {\em compressed sensing MRI} that exploits the implicit sparsity in MR images to significantly undersample the $k$-space and randomly select incoherent
(or complementary) samples regarding their spectral contribution to the MR image~\cite{Lustig07}.
This approach which does not require any multiple channel coil and thus remains useable with birdcage ones will not be addressed in this work.
The present paper is a contribution to parallel imaging
at conventional magnetic field strength~(e.g. 3 Tesla).
At 3 Tesla, Parallel Magnetic Resonance Imaging~(pMRI) is mainly useful in reception, i.e. at the
image \emph{reconstruction} step using multiple channel coils, while
higher magnetic fields (i.e. 7 Tesla) require parallel imaging
at the \emph{transmission} step as well~\cite{Katscher_03,Boulant09}.
Many methods like GRAPPA~(Generalized Autocalibrating Partially Parallel Acquisitions)~\cite{griswold_02}
and SENSE~(Sensitivity Encoding)~\cite{pruessmann_99} have been proposed in the literature
to reconstruct a full FOV image from multiple $k$-space under-sampled images acquired on separate channels.
The main difference between these two classes of methods lies in the space on which they operate:
GRAPPA performs multichannel full FOV reconstruction in the $k$-space whereas SENSE
carries out the unfolding process in the image domain: all the under-sampled images are first reconstructed
by inverse Fourier transform before combining them to unwrap the full FOV image.
Another difference is that GRAPPA is autocalibrated, while SENSE needs a separate coil sensitivity
estimation step based on a reference scan. Note however that an autocalibrated version of SENSE is also available for instance
in Siemens scanners and called \texttt{mSENSE}~hereafter.
SENSE as well as GRAPPA methods may suffer from strong artifacts when high values of
acceleration factors $R$ are considered in the imaging setup or when they are applied to
Echo Planar Imaging~(EPI), the sequence involved in fMRI experiments. These artifacts can drastically
disturb subsequent statistical analysis such as brain activation detection.
Regularized SENSE methods have been proposed in the literature to improve the robustness of the
solution~\cite{Liang_02,Ying_L_04,Liu_08_1,chaari_08,Liu_08_2}. Some of them apply quadratic or Total Variation~(TV)
regularizations while others resort to regularization in the wavelet transform domain~(e.g. UWR-SENSE~\cite{Chaari_MEDIA_2011}).
The latter strategy has proved its efficiency on the reconstruction of anatomical or functional~(resting-state only) data~\cite{chaari_08,Chaari_MEDIA_2011}.
More recently, UWR-SENSE has been assessed on EPI images and compared with \texttt{mSENSE}~on the same data acquired during a brain activation fMRI experiment~\cite{Chaari10e}.
This comparison was performed at the subject level on a few subjects only.
Besides, most of the available reconstruction methods in the literature operate slice by slice and thus reconstruct each slice irrespective of its neighbors.
Iterating over slices is thus necessary to get the whole 3D volume. This observation led us to consider 3D or whole brain reconstruction as a single
step in which all slices are treated together by making use of 3D wavelet transforms and 3D sparsity promoting regularization terms in the wavelet domain.
Following the same principle, an fMRI run usually consists of several hundred successive scans that are reconstructed independently one to another
and thus implicitly assumed to be independent of each other. Iterating over all acquired 3D volumes remains the classical approach to reconstruct the 4D or 3D~+~$t$ dataset
associated with an fMRI run. However, it has been shown for a long while that fMRI data are serially correlated in time
even under the null hypothesis~(i.e., ongoing activity only)~\cite{Aguirre97,Zarahn97,Purdon98}. To capture this dependence between successive time points, an autoregressive model has demonstrated its relevance~\cite{Woolrich01,Worsley02,Penny03}, especially
when its parameters and optimal order can vary in space for instance with tissue type~\cite{Woolrich04,Penny07}.
Hence, it makes sense to account for this temporal structure at the reconstruction step.
These two key ideas play a central role in the present
paper by extending the UWR-SENSE approach~\cite{Chaari_MEDIA_2011} through a fidelity data term combining all time points, which relies on a 3D wavelet transform and
an additional regularization term along the temporal dimension of the 4D dataset in the image domain. The development of the proposed
method (named 4D-UWR-SENSE) was made possible due to recent
advances in nonsmooth convex optimization. Indeed, it is based on a Parallel
ProXimal Algorithm~(PPXA) which is different from the ones employed in \cite{Chaari_MEDIA_2011}.
4D-UWR-SENSE leads to improvements in the retrieval of a reliable Signal-to-Noise Ratio~(SNR) between the acquired volumes and also in the enhancement
of the detection of BOLD effects or evoked activations that occur in response to the delivered stimuli in the fMRI experiment.
The present paper therefore aims at demonstrating that the 4D-UWR-SENSE approach outperfoms its SENSE-like alternatives not only in terms of reduced artifacts,
but also in terms of statistical sensitivity at the voxel and cluster levels, in intra-subject and group studies.
The rest of this paper is organized as follows. Section~\ref{sec:parallel} recalls the general parallel MRI framework. We then describe the proposed reconstruction
algorithm in Section~\ref{sec:algos} before illustrating related experimental results in Section~\ref{sec:validation}. Finally, some discussions and conclusions
are drawn in Section~\ref{sec:colclusion}.
\vspace*{-0.5cm}
\section{Parallel imaging in MRI}
\label{sec:parallel}
In parallel MRI, an array of $L$ coils is employed to measure the spin density $\overline{\rho}$ into the
object under investigation\footnote{The overbar is used to distinguish the ``true'' data from a generic
variable.}. The signal $\widetilde{d}_\ell$ received by each coil $\ell$ ($1\leq \ell \leq L$) is
the Fourier transform of the desired 2D field $\overline{\rho}\in\ensuremath{\mathbb R}^{X\times Y}$ on the
specified FOV weighted by the coil sensitivity profile $s_\ell$,
evaluated at some location $\vect{k}_r=(k_x,k_y)^{\mbox{\tiny\textsf{T}}}$ in
the $k$-space:
\begin{equation}
\widetilde{d}_\ell(\vect{k}_r)=\int\overline{\rho}(\vect{r})s_\ell(\vect{r})e^{-\imath 2\pi
\vect{k}^{\mbox{\tiny\textsf{T}}}_r\vect{r}}\,d\vect{r} +\widetilde{n}_\ell(\vect{k}_r), \label{eq:signal}
\end{equation}
\noindent where $\widetilde{n}_\ell(\vect{k}_r)$ is a coil-dependent
additive zero-mean Gaussian noise,
which is independent and identically distributed~(iid) in the $k$-space,
and $\vect{r}=(x,y)^{\mbox{\tiny\textsf{T}}}\in X\times Y$ is the spatial position in the
image domain ($\cdot^{\mbox{\tiny\textsf{T}}}$ being the transpose operator). The size of the reduced FOV acquired data $\widetilde{d}_\ell$ in the $k$-space clearly depends
on the sampling scheme. In contrast with cardiac imaging where the heart motion is significant during scanning,
a Cartesian coordinate system is generally adopted in the neuroimaging context.
In parallel MRI, the sampling period along the phase encoding direction
is $R$ times larger than the one used for conventional acquisition, $R \leq L$ being the reduction factor.
To recover full FOV images, many algorithms have been proposed but only SENSE-like~\cite{Pruessmann_K_99}
and GRAPPA-like~\cite{griswold_02} methods are provided by scanner manufacturers.
In what follows, we focus on SENSE-like methods operating in the spatial domain.
Let $\Delta y=\frac{Y}{R}$ be the aliasing period and $y$ the position in the image domain
along the phase encoding direction. Let $x$ be the position in the image domain along the frequency encoding
direction. A 2D inverse Fourier transform allows us to recover the measured signal in the spatial
domain. By accounting for the $k$-space undersampling at $R$-rate, the inverse Fourier transform
gives us the spatial counterpart of Eq.~\eqref{eq:signal} in matrix form:
\begin{align}\label{eq:matriciel}
\vect{d}(\vect{r}) &= \vect{S}(\vect{r}) \overline{\vect{\rho}}(\vect{r}) + \vect{n}(\vect{r}),
\end{align}
where
\begin{align}
\vect{S}(\vect{r})\,\mbox{$\buildrel\triangle\over =$}\,\left[
\begin{array}{ccc}
s_1(x,y)&\ldots&s_1(x,y+(R-1)\Delta y)\\
\vdots&\vdots&\vdots\\
s_L(x,y)&\ldots&s_L(x,y+(R-1)\Delta y)\\
\end{array}
\right]
, \quad&
\vect{n}(\vect{r})\,\mbox{$\buildrel\triangle\over =$}\,\left[
\begin{array}{c}
n_1(x,y)\\
n_2(x,y)\\
\vdots\\
n_L(x,y)\\
\end{array}
\right]
\nonumber
\end{align}
\begin{align}
\label{eq:defvrho}
& \overline{\vect{\rho}}(\vect{r})\,\mbox{$\buildrel\triangle\over =$}\,
\left[
\begin{array}{c}
\overline{\rho}(x,y)\\
\overline{\rho}(x,y+\Delta y)\\
\vdots\\
\overline{\rho}(x,y+(R-1)\Delta y)\\
\end{array}
\right] \quad \mathrm{and}
\quad
\vect{d}(\vect{r})\,\mbox{$\buildrel\triangle\over =$}\,
\left[
\begin{array}{c}
d_1(x,y)\\
d_2(x,y)\\
\vdots\\
d_L(x,y)\\
\end{array}
\right].
\end{align}
Based upon this model, the reconstruction step consists of solving Eq.~\eqref{eq:matriciel} so as to
recover $\overline{\vect{\rho}}(\vect{r})$ from $\vect{d}(\vect{r})$ and an estimate of
$\overline{\vect{S}}(\vect{r})$ at each spatial position $\vect{r}=(x,y)^{\mbox{\tiny\textsf{T}}}$. The spatial mixture
or \emph{sensitivity} matrix $\overline{\vect{S}}(\vect{r})$ is estimated using a reference scan and varies
according to the coil geometry. Note that the coil images $(d_\ell)_{1\leq l \leq L}$ as well as the
sought image $\overline{\rho}$ are complex-valued, although $|\overline{\rho}|$ is only considered for
visualization purposes. The next section describes the widely used SENSE algorithm as well as its
regularized extensions.
\vspace*{-0.5cm}
\section{Reconstruction algorithms} \label{sec:algos}
\subsection{1D-SENSE}
In its simplest form, SENSE imaging amounts to solving a one-dimensional inversion problem
due to the separability of the Fourier transform. Note however that this inverse problem admits a
two-dimensional extension in 3D imaging sequences like Echo Volume Imaging~(EVI)~\cite{Rabrait07} where
undersampling occurs in two $k$-space directions. The 1D-SENSE
reconstruction method~\cite{Pruessmann_K_99} actually minimizes a Weighted Least Squares~(WLS) criterion
$\mathcal{J}_{\rm WLS}$ given by:
\begin{equation}
\label{eq:crit_WLS}
\mathcal{J}_{\rm WLS}(\rho) = \sum_{\mathbf{r}\in \{1,\ldots,X\} \times \{1,\ldots,Y/R\}} \parallel \vect{d}(\vect{r})-\vect{S}(\vect{r})\vect{\rho}(\vect{r}) \parallel^2_{\vect{\Psi}^{-1}},
\end{equation}
\noindent where $\|\cdot\|_{\vect{\Psi}^{-1}}= \sqrt{(\cdot)^{\mbox{\tiny\textsf{H}}}\vect{\Psi}^{-1}(\cdot)}$, and the noise covariance matrix $\vect{\Psi}$ is usually estimated based on $L$ acquired images
$(\underbar{d}_{\ell})_{1\leq \ell \leq L}$ from all coils without radio frequency pulse.
Hence, the SENSE full FOV image is nothing but
the maximum likelihood estimate under Gaussian noise assumption, which admits the following closed-form expression
at each spatial position $\vect{r}$:
\begin{equation}
\widehat{\vect{\rho}}_{\rm WLS}(\vect{r}) = \pth{\vect{S}^{\mbox{\tiny\textsf{H}}}(\vect{r})\vect{\Psi}^{-1}\vect{S}(\vect{r})}^{\sharp}\vect{S}^{\mbox{\tiny\textsf{H}}}(\vect{r})
\vect{\Psi}^{-1}\vect{d}(\vect{r}),
\end{equation}
\noindent where
$(\cdot)^{\mbox{\tiny\textsf{H}}}$~(resp. $(\cdot)^{\sharp}$) stands for the transposed complex conjugate~(resp.
pseudo-inverse).
It should be noticed here that the described 1D-SENSE reconstruction method has been designed to reconstruct one slice~(2D image).
To reconstruct a full volume, the 1D-SENSE reconstruction algorithm has to be iterated over all slices.\\
In practice, the performance of the 1D-SENSE method is limited because of {\em i)} the presence of
distortions in the measurements
$\vect{d}(\vect{r})$, {\em ii)} the ill-conditioning of $\vect{S}(\vect{r})$, in particular
at locations $\vect{r}$ close to the image center and {\em iii)} the presence of errors in the
estimation of $\vect{S}(\vect{r})$ mainly at brain/air interfaces. To enhance the robustness of the
solution to this ill-posed problem, a regularization is usually introduced in the
reconstruction process. To improve results obtained with quadratic regularization techniques \cite{Liang_02,Ying_L_04},
edge-preserving regularization has been widely investigated in the pMRI reconstruction literature. For instance, reconstruction methods based on
Total Variation~(TV)
regularization have been proposed in a number of recent works like \cite{keeling_03,Liu_08}. However, TV is mostly adapted to piecewise
constant images, which are not always acurate models in MRI, especially in fMRI.
As investigated by \textit{Chaari et al.}~\cite{Chaari_MEDIA_2011} and \textit{Liu et al.}~\cite{Liu_08_2},
regularization in the Wavelet Transform (WT) domain is a powerful tool to improve SENSE reconstruction.
In what follows, we summarize the principles of the wavelet-based regularization approach.
\subsection{Proposed wavelet Regularized-SENSE}\label{subsec:WRSENSE}
Akin to \cite{Chaari_MEDIA_2011} where a regularized reconstruction algorithm relying on 2D separable WTs was investigated, to the best of our knowledge, all the existing approaches
in the pMRI regularization literature proceed slice by slice. The drawback of this strategy is that no spatial continuity between
adjacent slices is taken into account since the slices are processed independently.
Moreover, since the whole brain volume has to be acquired several times in an fMRI study,
iterating over all the acquired 3D volumes is then necessary in order to reconstruct a 4D data volume corresponding to an fMRI session.
Consequently, the 3D volumes are supposed independent whereas fMRI time-series are serially correlated in time because of two distinct effects:
the BOLD signal itself is a low-pass filtered version of the neural activity, and physiological artifacts make the fMRI time points strongly dependent.
For these reasons, modeling temporal dependence across scans at the reconstruction step may impact subsequent statistical analysis.
This has motivated the extension of the wavelet regularized reconstruction approach in \cite{Chaari_MEDIA_2011} in order
to:
\begin{itemize}
\item account for 3D spatial dependencies between adjacent slices by using 3D WTs,
\item exploit the temporal dependency between acquired 3D volumes by applying an additional regularization term along the temporal dimension of the 4D dataset.
\end{itemize}
This additional regularization will help in increasing the Signal to Noise Ratio~(SNR) through the acquired volumes, and therefore enhance the reliability of
the statistical analysis in fMRI. These temporal dependencies have also been used in the dynamic MRI literature in order to improve the reconstruction
quality in conventional MRI \cite{Sumbul_2009}. However, since the imaged object geometry in the latter context generally changes during the acquisition,
taking into account the temporal regularization in the reconstruction
process is very difficult.\\
To deal with a 4D reconstruction of the $N_r$ acquired volumes, we will first rewrite the observation model in Eq.~\eqref{eq:matriciel} as follows:
\begin{equation}
\vect{d}^t(\vect{r}) = \vect{S}(\vect{r})\vect{\rho}^t(\vect{r}) + \vect{n}^t(\vect{r}),
\end{equation}
\noindent where $t \in \{1,\ldots,N_r\}$ is the acquisition time and $\vect{r} = (x,y,z)$ is the 3D spatial position, $z \in \{1, \ldots, Z\}$ being the position along
the third direction (slice selection one).\\
At a given time $t$, the full FOV 3D complex-valued image $\overline{\rho}^t$ of size $X \times Y \times Z$ can be
seen as an element of the Euclidean space $\mathbb{C}^K$ with $K = X \times Y \times Z$ endowed
with the standard inner product $\scal{\cdot}{\cdot}$ and norm $\|\cdot \|$.
We employ a dyadic 3D orthonormal wavelet decomposition operator $T$ over $j_\mathrm{max}$ resolution levels.
The coefficient field resulting from the wavelet
decomposition of a target image $\rho^t$ is defined as
$\zeta^t =\big({\sbmm{\zeta}}\XS^t_{a}, ({\sbmm{\zeta}}\XS^t_{o,j})_{o\in \mathbb{O},1 \le j \le j_\mathrm{max}}\big)$ with
$o \in \mathbb{O} =\{0,1\}^3\setminus \{(0,0,0)\}$, ${\sbmm{\zeta}}\XS^t_{a} = (\zeta^t_{a,k})_{1 \le k\le K_{j_\mathrm{max}}}$ and ${\sbmm{\zeta}}\XS^t_{o,j}=
(\zeta^t_{o,j,k})_{1\le k \le K_j}$
where $K_{j}= K2^{-3j}$ is the number of wavelet coefficients in a given subband at resolution $j$ (by assuming that $X$, $Y$ and $Z$ are multiple of
$2^{j_{\mathrm{max}}}$).
Adopting such a notation, the wavelet coefficients have been reindexed so that ${\sbmm{\zeta}}\XS^t_{a}$ denotes the approximation coefficient vector at the resolution level
$j_\mathrm{max}$, while
${\sbmm{\zeta}}\XS^t_{o,j}$ denotes the detail coefficient vector at the orientation $o$ and resolution level $j$.
Using 3D dyadic WTs allows us to smooth reconstruction artifacts along the slice selection direction, which is not possible using a
slice by slice operating approach.\\
The proposed regularization procedure relies on the introduction of two penalty terms. The first penalty
term describes the prior spatial knowledge about the wavelet coefficients of the target solution and it is expressed as:
\begin{equation}
g(\zeta) = \sum_{t = 1}^{N_r} \Big[\sum_{k=1}^{K_{j_\mathrm{max}}} \Phi_{a}(\zeta^t_{a,k}) + \sum_{o\in \mathbb{O}} \sum_{j=1}^{j_{\mathrm{max}}} \sum_{k=1}^{K_j}
\Phi_{o,j}(\zeta^t_{o,j,k}) \Big],
\end{equation}
where $\zeta = (\zeta^1,\zeta^2,\ldots,\zeta^{N_r})$ and we have, for every $o \in \mathbb{O}$ and $j \in \{1,\ldots,j_{\rm max}\}$,
\begin{equation}
\forall \xi \in {\ensuremath{\mathbb C}},\quad \; \Phi_{o,j}(\xi) = \Phi^{\rm Re}_{o,j}(\xi) + \Phi^{\rm Im}_{o,j}(\xi)
\end{equation}
\noindent where
$\Phi^{\rm Re}_{o,j}(\xi) = \alpha_{o,j}^{\mathrm{Re}}|\mathrm{Re}(\xi - \mu_{o,j})| + \frac{\beta_{o,j}^{\mathrm{Re}}}{2}|\mathrm{Re}(\xi - \mu_{o,j})|^2$ and
$\Phi^{\rm Im}_{o,j}(\xi) = \alpha_{o,j}^{\mathrm{Im}}|\mathrm{Im}(\xi - \mu_{o,j})| + \frac{\beta_{o,j}^{\mathrm{Im}}}{2}|\mathrm{Im}(\xi - \mu_{o,j})|^2$
with $\mu_{o,j} = \mu^{\rm Re}_{o,j} + \imath \mu^{\rm Im}_{o,j}\in {\ensuremath{\mathbb C}}$,
and $\alpha_{o,j}^{\mathrm{Re}}$, $\beta_{o,j}^{\mathrm{Re}}$,
$\alpha_{o,j}^{\mathrm{Im}}$, $\beta_{o,j}^{\mathrm{Im}}$ are some positive real
constants.
Hereabove, $\mathrm{Re}(\cdot)$ and $\mathrm{Im}(\cdot)$ (or $\cdot^{\mathrm{Re}}$ and
$\cdot^{\mathrm{Im}}$) stand for the real and imaginary parts, respectively.
A similar model is adopted for the approximation coefficients.
The second regularization term penalizes the temporal variation
between successive 3D volumes:
\begin{equation}
h(\zeta) = \kappa \sum_{t = 2}^{N_r} \Vert T^*\zeta^{t} - T^*\zeta^{t-1} \Vert_p^p
\end{equation}
\noindent where $T^*$ is the 3D wavelet reconstruction operator.
The prior parameters $\vect{\alpha}_{o,j} = (\alpha_{o,j}^{\mathrm{Re}},
\alpha_{o,j}^{\mathrm{Im}})$,
$\vect{\beta}_{o,j}=(\beta_{o,j}^{\mathrm{Re}},\beta_{o,j}^{\mathrm{Im}})$,
${\sbmm{\mu}}\XS_{o,j}=(\mu^{\mathrm{Re}}_{o,j},\mu^{\mathrm{Im}}_{o,j})$, $\kappa \in \ensuremath{\mathbb R}_+$ and $p \in [1,+\infty[$ are unknown and they need to be
estimated\linebreak (see~\ref{append:a2}).\\
The operator $T^*$ is then applied to each component $\zeta^t$ of $\zeta$ to obtain the reconstructed 3D volume $\rho^t$ related to the
acquisition time $t$.
It should be noticed here that other choices for the penalty functions are also possible provided that the convexity of the resulting optimality criterion is ensured.
This condition enables the use of fast and efficient convex optimization algorithms. Adopting this formulation, the minimization process plays a prominent role in the reconstruction process, as detailed in~\ref{append:a1}.
\section{Experimental validation in fMRI}\label{sec:results} \label{sec:validation}
This section is dedicated to the experimental validation of the
4D-UWR-SENSE reconstruction algorithm we proposed in Section~\ref{subsec:WRSENSE}.
Results of subject and group-level fMRI statistical analyses are
compared for two reconstruction pipelines: one available on the Siemens
workstation and our own pipeline involving for the sake of completeness either
the early UWR-SENSE~\cite{Chaari_MEDIA_2011} or the 4D-UWR-SENSE version of
the proposed pMRI reconstruction algorithm.
In what follows, we first describe the fMRI acquisition setup
and the experimental design.
\subsection{Experimental data} \label{subsec:realdata}
For validation purpose, we acquired fMRI data on a 3 T Siemens Trio magnet
using a Gradient-Echo EPI~(GE-EPI)
sequence ($TE=30~\rm ms$, $TR=2400~\rm ms$, slice thickness = 3~mm, transversal
orientation, FOV = $192~\mathrm{mm}^2$)
during a cognitive \textit{localizer} \cite{Pinel_07} protocol.
This experiment has been designed to map auditory, visual and motor
brain functions as well as higher cognitive tasks such as number processing and
language comprehension~(listening and reading). It consisted of a single session of $N_r = 128$ scans.
The paradigm was a fast event-related design comprising sixty auditory,
visual and motor stimuli, defined in ten experimental
conditions (auditory and visual sentences, auditory and visual calculations, left/right auditory and
visual clicks, horizontal and vertical checkerboards).
An $L = 32$ channel coil was used to enable parallel imaging.
Ethics approval was given by the local research ethics committee,
and fifteen subjects gave written informed consent for participation.
For each subject,
fMRI data were collected at the $2\times2~\mathrm{mm}^2$ spatial in-plane
resolution using different
reduction factors~($R = 2$ or $R = 4$). Based on the raw data files
delivered by the scanner, reduced FOV EPI images were reconstructed
as detailed in Fig.~\ref{fig:reading_data}. This reconstruction requires
two specific treatments:
\begin{itemize}
\item[\textit{i)}] \textit{ $k$-space regridding} to account for the
non-uniform $k$-space sampling during readout gradient ramp,
which occurs in fast MRI sequences like GE-EPI;
\item[\textit{ii)}] \textit{Nyquist ghosting correction} to remove the odd-even
echo inconsistencies during $k$-space acquisition of EPI images
\end{itemize}
\begin{figure}[!ht]
\centering
\scalebox{0.46}[0.5]{\input{Figures/reading_data2.pstex_t}}\vspace*{-.3cm}
\caption{Reconstruction pipeline of reduced FOV EPI images
from the raw FID data.\label{fig:reading_data}}
\end{figure}
\noindent Once the reduced FOV images are available, the proposed pMRI
4D-UWR-SENSE algorithm and its early UWR-SENSE version have been
utilized in a final step to reconstruct the full FOV EPI
images and compared to the \texttt{mSENSE}\footnote{SENSE reconstruction implemented
by the Siemens scanner} Siemens solution.
For the wavelet-based regularization, dyadic~($M = 2$) \textit{Symmlet} orthonormal wavelet bases~\cite{daubechies_92}
associated with filters of length 8 have been used over $j_{\mathrm{max}}=3$ resolution levels.
The reconstructed EPI images then enter in our fMRI study in order to measure the impact of
the reconstructor choice on brain activity detection.
Note also that the proposed reconstruction algorithm
requires the estimation of the coil sensitivity maps~(matrix ${\sbm{S}}\XS} \def\sb{{\sbm{s}}\XS(\cdot)$ in
Eq.~\eqref{eq:matriciel}). As proposed in~\cite{pruessmann_99},
the latter were estimated by dividing the coil-specific images by the module of the
Sum Of Squares~(SOS) images, which are computed from the specific acquisition of
the $k$-space center~(24 lines) before the $N_r$ scans.
Fig.~\ref{fig:slice_Axial} compares the two pMRI reconstruction algorithms
to illustrate on axial, coronal and sagittal slices how the \texttt{mSENSE}~reconstruction
artifacts have been removed using the 4D-UWR-SENSE approach.
The \texttt{mSENSE}~reconstructed images acutally present large artifacts located both
at the center and boundaries of the brain in sensory and cognitive
regions~(temporal lobes, frontal and motor cortices, ...). This
results in SNR loss and thus may have a dramatic impact for activation detection
in these brain regions. Note that these conclusions are reproducible across
subjects although
the artifacts may appear on different slices~(see red circles in
Fig.\ref{fig:slice_Axial}). It is also worth mentioning that
in contrast to the Siemens reconstructor our pipeline
does not involve any spatial filtering step to improve signal homogeneity
across the brain: this explains why the images shown in Fig.~\ref{fig:slice_Axial}
and delivered by our 4D-UWR-SENSE algorithm seem less homogeneous.
However, bias field correction can be applied if necessary using specific tools
such as those available in BrainVISA\footnote{\url{http://brainvisa.info}.}.
\begin{figure}[!ht]
\centering
\begin{tabular}{c| c c c c c c}
&&\texttt{mSENSE}&4D-UWR-SENSE\\
&\raisebox{1.1cm}{Axial}&
\includegraphics[width=2.2cm, height=2cm]{mSENSE_R2_Axial_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R2_Axial_MF090124.eps}\\
\raisebox{1.1cm}{$R=2$}&
\raisebox{1.1cm}{Coronal}&\includegraphics[width=2.2cm, height=2cm]{mSENSE_R2_Coronal_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R2_Coronal_MF090124.eps}\\
&\raisebox{1.1cm}{Sagittal}&\includegraphics[width=2.2cm, height=2cm]{mSENSE_R2_Sagittal_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R2_Sagittal_MF090124.eps}\\
\hline
&\raisebox{1.1cm}{Axial}&\includegraphics[width=2.2cm, height=2cm]{mSENSE_R4_Axial_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R4_Axial_MF090124.eps}\\
\raisebox{1.1cm}{$R=4$}&
\raisebox{1.1cm}{Coronal}&
\includegraphics[width=2.2cm, height=2cm]{mSENSE_R4_Coronal_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R4_Coronal_MF090124.eps}\\
&\raisebox{1.1cm}{Sagittal}&
\includegraphics[width=2.2cm, height=2cm]{mSENSE_R4_Sagittal_MF090124.eps}&
\includegraphics[width=2.2cm, height=2cm]{4D_R4_Sagittal_MF090124.eps}
\end{tabular
\caption{\textbf{Axial}, \textbf{Coronal} and \textbf{Sagittal} reconstructed slices using \texttt{mSENSE}~and 4D-UWR-SENSE for $R=2$ and $R=4$ with
$2\times 2~\rm{mm}^2$ in-plane spatial resolution. Red circles and ellipsoids indicate the position
of reconstruction artifacts using \texttt{mSENSE}.
\label{fig:slice_Axial}}
\end{figure}
Regarding the computation burden, the \texttt{mSENSE}~algorithm is carried out on-line and remains
compatible with real time processing. On the other hand, our pipeline is carried out off-line
and requires much more computation. For illustration purpose, on a biprocessor quadicore {\tt Intel
Xeon CPU}@~2.67GHz, one EPI slice is reconstructed in 4~s using the UWR-SENSE algorithm.
Using parallel computing strategy and multithreading~(through the \texttt{OMP} library), each EPI volume
consisting of 40 slices is reconstruced in 22~s. This makes the whole series of 128 EPI images available
in about 47~min.
In contrast, the proposed 4D-UWR-SENSE achieves the reconstruction of the series in about 40 min, but requires larger memory space due to large data volume
processed simultaneously.
\subsection{fMRI data pre-processings}\label{subsec:Preprocs}
Irrespective of the reconstruction pipeline,
the full FOV fMRI images were then preprocessed using the SPM5
software\footnote{\url{http://www.fil.ion.ucl.ac.uk/spm/software/spm5/}}:
preprocessing involves realignment, correction for motion and differences in
slice acquisition time, spatial normalization, and smoothing with an isotropic
Gaussian kernel of $4$mm full-width at half-maximum.
Anatomical normalization to MNI space was performed by coregistration of the
functional images with the anatomical T1 scan acquired with the
thirty two channel-head coil. Parameters for the normalization
to MNI space were estimated by normalizing this scan to the T1 MNI
template provided by SPM5, and were subsequently applied to all functional images.
\subsection{Subject-level analysis}\label{subsec:Subject-level-analysis}
A General Linear Model~(GLM) was constructed to capture stimulus-related
BOLD response.
As shown in Fig.~\ref{fig:design}, the design matrix relies on ten
experimental conditions and thus made up of twenty one regressors
corresponding to stick functions convolved with
the canonical Haemodynamic Response Function (HRF) and its
first temporal derivative, the last regressor modelling the baseline.
This GLM was then fitted to the same acquired images but reconstructed
using either the Siemens reconstructor or our own pipeline, which in the following
is derived from the early UWR-SENSE method~\cite{Chaari_MEDIA_2011} and
from its 4D-UWR-SENSE extension we propose here.
\begin{figure}[!ht]
\centering
\includegraphics[width=8cm, height=6cm]{design.eps}
\caption{(a): design matrix and the \texttt{Lc-Rc}\XS contrast involving two
conditions~(grouping auditory and visual modalities);
(b): design matrix and the aC-aS\XS contrast involving four conditions~(sentence,
computation, left click, right click).\label{fig:design}}
\end{figure}
Here, contrast estimate images for motor responses and higher cognitive
functions~(computation, language) were subjected to further analyses at the
subject and group levels. These two contrast are complementary since the expected activations
lie in different brain regions and thus can be differentially corrupted by
reconstruction artifacts as outlined in Fig.~\ref{fig:slice_Axial}.
More precisely, we studied:
\begin{itemize}
\item the {\bf Auditory computation vs. Auditory sentence}~(aC-aS\XS) contrast which is
supposed to elicit evoked activity in the frontal and parietal lobes, since
solving mental arithmetic task involves working memory and more specifically
the intra-parietal sulcus~\cite{Dehaene99}: see Fig.~\ref{fig:design}(b);
\item the {\bf Left click vs. Right click}~(\texttt{Lc-Rc}\XS) contrast for which we expect
evoked activity in the right motor cortex~(precentral gyrus, middle frontal gyrus).
Indeed, the \texttt{Lc-Rc}\XS contrast defines a
compound comparison which involves two motor stimuli which are presented either in the
visual or auditory modality. This comparison aims therefore at detecting
lateralization effect in the motor cortex: see Fig.~\ref{fig:design}(a).
\end{itemize}
Interestingly, these two contrasts were chosen because they summarized well
different situations~(large vs small activation clusters, distributed vs focal
activation pattern, bilateral vs unilateral activity) that occurred for this paradigm when looking at sensory
areas~(visual, auditory, motor) or regions involved in higher cognitive
functions~(reading, calculation). In the following,
our results are reported in terms of Student-$t$ maps
thresholded at a cluster-level $p=0.05$ corrected for multiple
comparisons according to the FamilyWise Error Rate~(FWER)~\cite{Nichols03,Brett_04}.
Complementary statistical tables provide corrected cluster and voxel-level $p$-values,
maximal $t$-scores and corresponding peak positions both for $R=2$ and $R=4$.
Note that clusters are listed in a decreasing order of significance.
Concerning the aC-aS\XS contrast, Fig.~\ref{fig:res_T_A-V}[top] shows for the most significant slice
and $R=2$ that all pMRI reconstruction algorithms succeed in finding evoked activity
in the left parietal and frontal cortices, more precisely in the inferior
parietal lobule and middle frontal gyrus according to the AAL template\footnote{available
in the \texttt{xjView} toolbox of SPM5.}.
Table~\ref{tab:StatRes2all} also confirms
a bilateral activity pattern in parietal regions for $R=2$.
Moreover, for $R=4$ Fig.~\ref{fig:res_T_A-V}[bottom] illustrates
that our pipeline~(UWR-SENSE and 4D-UWR-SENSE) and preferentially
the proposed 4D-UWR-SENSE scheme enables to retrieve reliable
frontal activity elicited by mental calculation, which is lost by the
the \texttt{mSENSE}~algorithm.
From a quantitative viewpoint, the proposed 4D-UWR-SENSE algorithm finds
larger clusters whose local maxima are more significant
than the ones obtained using \texttt{mSENSE}~and UWR-SENSE, as reported in
Table~\ref{tab:StatRes2all}.
Concerning the most significant cluster for $R=2$, the peak positions
remain stable whatever the reconstruction algorithm.
However, examining their significance level,
one can first measure the benefits of wavelet-based regularization
when comparing UWR-SENSE with \texttt{mSENSE}~results and then additional
positive effects of temporal regularization and 3D wavelet decomposition
when looking at the 4D-UWR-SENSE results. These benefits are also demonstrated for $R=4$.
\begin{figure}[!ht
\centering
\begin{tabular}{c c c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=2$}&\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R2_CS_LE.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R2_CS_LE.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R2_CS_LE.eps}\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=4$}&\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R4_CS_LE.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R4_CS_LE.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R4_CS_LE.eps}
\end{tabular}\vspace*{-.3cm}
\caption{Subject-level student-$t$ maps superimposed to anatomical MRI for the aC-aS\XS contrast.
Data have been reconstructed using the~\texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE, respectively.
Neurological convention: {\bf left is left}.
The blue cross shows the maximum activation peak. \label{fig:res_T_A-V}}
\end{figure}
\begin{table}[!ht]
\centering
\caption{Significant statistical results at the subject-level for the aC-aS\XS contrast (corrected for multiple comparisons at $p=0.05$).
Images were reconstructed using the~\texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE algorithm for $R=2$ and $R=4$.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\cline{3-7}
\cline{3-7}
\multicolumn{2}{c}{}&\multicolumn{2}{|c|}{cluster-level}&\multicolumn{3}{|c|}{voxel-level}\\
\cline{3-7}
\multicolumn{2}{c|}{}&p-value&Size&p-value&T-score& Position\\
\hline
&\multirow{4}{*}{\texttt{mSENSE}} & $< 10^{-3}$ &320 &$< 10^{-3}$ & 6.40&-32 -76 45 \\
\cline{3-7}
\multirow{12}{*}{$R=2$} & & $< 10^{-3}$ &163 &$< 10^{-3}$ & 5.96&-4 -70 54 \\
\cline{3-7}
& &$< 10^{-3}$ &121 &$< 10^{-3}$ & 6.34&34 -74 39 \\
\cline{3-7}
&&$< 10^{-3}$ &94 &$< 10^{-3}$ & 6.83&-38 4 24 \\
\cline{2-7}
\cline{2-7}
&\multirow{4}{*}{UWR-SENSE} & $< 10^{-3}$ &407 &$< 10^{-3}$& 6.59&-32 -76 45 \\
\cline{3-7}
& & $< 10^{-3}$ &164 &$< 10^{-3}$& 5.69&-6 -70 54 \\
\cline{3-7}
& & $< 10^{-3}$ &159 &$< 10^{-3}$& 5.84&32 -70 39 \\
\cline{3-7}
& & $< 10^{-3}$ &155 &$< 10^{-3}$& 6.87&-44 4 24 \\
\cline{2-7}
\cline{2-7}
&\multirow{4}{*}{4D-UWR-SENSE} &$< 10^{-3}$ &\textbf{454} &$< 10^{-3}$ & 6.54& -32 -76 45 \\
\cline{3-7}
& & $< 10^{-3}$ &199& $< 10^{-3}$ & 5.43& -6 26 21 \\
\cline{3-7}
& &$< 10^{-3}$ &183 & $< 10^{-3}$ & 5.89& 32 -70 39 \\
\cline{3-7}
& &$< 10^{-3}$ &170 & $< 10^{-3}$ &\textbf{ 6.90}& -44 4 24 \\
\hline
\hline
\multirow{6}{*}{$R=4$}&\multicolumn{1}{|c|}{\texttt{mSENSE}} &$< 10^{-3}$ & 58& 0.028& 5.16&-30 -72 48\\
\cline{2-7}
\cline{2-7}
&\multirow{2}{*}{4D-UWR-SENSE} &$< 10^{-3}$ & 94& 0.003& 5.91&-32 -70 48\\
\cline{3-7}
& &$< 10^{-3}$ & 60& 0.044& 4.42&-6 -72 54\\
\cline{2-7}
\cline{2-7}
&\multirow{3}{*}{4D-UWR-SENSE} &$< 10^{-3}$ & \textbf{152} &$< 10^{-3}$&\textbf{6.36}&-32 -70 48 \\
\cline{3-7}
& &$< 10^{-3}$ & 36 &0.009&5.01&-4 -78 48 \\
\cline{3-7}
& &$< 10^{-3}$ & 29 &0.004&5.30&-34 6 27 \\
\hline
\end{tabular}
\label{tab:StatRes2all}
\end{table}
Fig.~\ref{fig:ACAS_variability} illustrates another property of the
proposed pMRI pipeline, ie its robustness to the between-subject variability.
Indeed, when comparing subject-level student-$t$ maps
reconstructed using the different pipelines~($R=2$), it can
be observed that the \texttt{mSENSE}~algorithm
fails to detect any activation cluster in the expected regions
for the second subject~(see Fig.~\ref{fig:ACAS_variability}[bottom]). In contrast,
our 4D-UWR-SENSE method retrieves more coherent activity
while not exactly at the same position as for the first subject.
\begin{figure}[!ht]
\centering
\begin{tabular}{cc c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.1cm}\raisebox{1.8cm}{\footnotesize \texttt{Subj.~1}}&\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R2_CS_LE.eps}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R2_CS_LE}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R2_CS_LE}\\
\hspace*{-0.1cm}\raisebox{1.8cm}{\footnotesize \texttt{Subj.~2}}&\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R2_CS.eps}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R2_CS}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R2_CS}
\end{tabular}
\caption{Between-subject variability of detected activation for the aC-aS\XS contrast at $R=2$.
Neurological convention. The blue cross shows the activation peak.\label{fig:ACAS_variability}}
\end{figure}
\begin{figure}[!ht]
\centering
\begin{tabular}{c c c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=2$}&
\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR2_mSENSE}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR2_UWRSENSE_2D}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR2_UWRSENSE_4D}\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=4$}&
\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_mSENSE}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_UWRSENSE_2D}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_UWRSENSE_4D}
\end{tabular}\vspace*{-.3cm}
\caption{Subject-level student-$t$ maps superimposed to anatomical MRI for the \texttt{Lc-Rc}\XS
contrast. Data have been reconstructed using the \texttt{mSENSE}, UWR-SENSE
and 4D-UWR-SENSE, respectively. Neurological convention.
The blue cross shows the activation peak. \label{fig:res_T_Lc-Rc}}
\end{figure}
For the \texttt{Lc-Rc}\XS contrast on the data acquired with $R=2$,
Fig.~\ref{fig:res_T_Lc-Rc}[top] shows that all
reconstruction methods enable to retrieve expected activation in the right
precentral gyrus. However, when looking more carefully at the statistical
results~(see Table~\ref{tab:StatRes2allRcLc}), our pipeline
and more preferentially the 4D-UWR-SENSE algorithm retrieves an additional cluster
in the right middle frontal gyrus. On data acquired with $R=4$,
the same \texttt{Lc-Rc}\XS contrast elicits similar activations, ie in the same region.
As demonstrated in Fig.~\ref{fig:res_T_Lc-Rc}[bottom], this activity is enhanced
when pMRI reconstruction is performed with our pipeline.
Quantitative results in Table~\ref{tab:StatRes2allRcLc} confirms numerically
what can be observed in Fig.~\ref{fig:res_T_Lc-Rc}:
larger clusters with higher local $t$-scores
are detected using the 4D-UWR-SENSE algorithm, both for $R=2$ and $R=4$.
Also, a larger number of clusters is retrieved for $R=2$ using wavelet-based regularization.
\begin{table}[!ht]
\centering
\caption{Significant statistical results at the subject-level for the \texttt{Lc-Rc}\XS contrast (corrected for multiple comparisons at $p=0.05$).
Images were reconstructed using the \texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE algorithms for $R=2$ and $R=4$.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\cline{3-7}
\cline{3-7}
\multicolumn{2}{c}{}&\multicolumn{2}{|c|}{cluster-level}&\multicolumn{3}{|c|}{voxel-level}\\
\cline{3-7}
\multicolumn{2}{c|}{}&p-value&Size&p-value&T-score& Position\\
\hline
\multirow{5}{*}{$R=2$}&
\multicolumn{1}{|c|}{\texttt{mSENSE}} &$ <10^{-3}$ &79&$ <10^{-3}$ & 6.49&38 -26 66 \\
\cline{2-7}
\cline{2-7}
&\multirow{2}{*}{UWR-SENSE} &$ <10^{-3}$ &144&0.004& 5.82& 40 -22 63 \\
\cline{3-7}
& &$0.03$ &21&0.064& 4.19& 24 -8 63 \\
\cline{2-7}
\cline{2-7}
&\multirow{2}{*}{4D-UWR-SENSE} &$<10^{-3}$ &\textbf{172}&0.001&\textbf{ 6.78}& 34 -24 69 \\
\cline{3-7}
& &$<10^{-3}$ &79&0.001&6.49& 38 -26 66 \\
\hline
\hline
\multirow{3}{*}{$R=4$}&\multicolumn{1}{|c|}{\texttt{mSENSE}} &0.006 & 21 & 0.295&4.82&34 -28 63\\
\cline{2-7}
\cline{2-7}
&\multicolumn{1}{|c|}{UWR-SENSE} &$< 10^{-3}$ & 33& 0.120& 5.06&40 -24 66\\
\cline{2-7}
\cline{2-7}
&\multicolumn{1}{|c|}{4D-UWR-SENSE} &$< 10^{-3}$&\textbf{51}& 0.006&\textbf{5.57}&40 -24 66\\
\hline
\end{tabular}
\label{tab:StatRes2allRcLc}
\end{table}
Fig.~\ref{fig:LcRc_variability} reports on the robustness of the proposed
pMRI pipeline to the between-subject variability for this motor contrast.
Since sensory functions are expected to generate larger BOLD effects~(higher SNR)
and appears more stable, our comparison takes place at $R=4$.
Two subject-level student-$t$ maps reconstructed using the different pMRI algorithms
are compared: in Fig.~\ref{fig:LcRc_variability}. For the second subject,
one can observe that the \texttt{mSENSE}~algorithm fails to detect any activation cluster in
the right motor cortex. In contrast, our 4D-UWR-SENSE method retrieves
more coherent activity for this second subject in the expected region.
\begin{figure}[!ht]
\centering
\begin{tabular}{cc c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.1cm}\raisebox{1.8cm}{\footnotesize \texttt{Subj.~1}}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_mSENSE}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_UWRSENSE_2D}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{Lc-Rc_MF_acqR4_UWRSENSE_4D}\\
\hspace*{-0.1cm}\raisebox{1.8cm}{\footnotesize \texttt{Subj.~5}}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{AD080125_Lc-Rc_mSENSE_R4}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{AD080125_Lc-Rc_2DUWRSENSE_R4}&
\hspace*{-0.35cm}\includegraphics[width=3.7cm, height=3.7cm]{AD080125_Lc-Rc_4DUWRSENSE_R4}
\end{tabular}
\caption{Between-subject variability of detected activation for the \texttt{Lc-Rc}\XS contrast at $R=4$.
Neurological convention. The blue cross shows the activation peak.\label{fig:LcRc_variability}}
\end{figure}
To summarize, on these two contrasts our 4D-UWR-SENSE algorithm always
outperforms the alternative reconstruction methods
in terms of statistical significance~(number of clusters, cluster extent,
peak values,...) but also in terms of robustness.
\subsection{Group-level analysis}
Due to between-subject anatomical and functional variability,
group-level analysis is necessary in order to derive robust and reproducible conclusions
at the population level. For this validation, random effect analyses~(RFX) involving fifteen
healthy subjects have been conducted on the contrast maps we previously investigated at
the subject level. More precisely, one-sample Student-$t$ test was performed
on the subject-level contrast images~(eg, \texttt{Lc-Rc}\XS, aC-aS\XS,... images) using SPM5.
\begin{figure}[!ht]
\centering
\begin{tabular}{c c c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=2$}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R2_CS_GR.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R2_CS_GR.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R2_CS_GR.eps}\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=4$}&
\includegraphics[width=3.7cm, height=3.7cm]{ac-as_R4_RFX_mSENSE.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{ac-as_R4_RFX_UWRSENSE-2D.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{ac-as_R4_RFX_UWRSENSE-4D.eps}
\end{tabular}
\caption{Group-level student-$t$ maps for the aC-aS\XS contrast where data have been reconstructed using the \texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE for $R=2$ and $R=4$.
Neurological convention. Red arrows indicate the global maximum activation peak.\label{fig:res_G_A-V}}
\end{figure}
For the aC-aS\XS contrast, Maximum Intensity Projection~(MIP) student-$t$ maps
are shown in Fig.~\ref{fig:res_G_A-V}. First, they illustrate that
irrespective of the reconstruction method larger and more significant
activations are found on datasets acquired with $R=2$ given the better SNR.
Second, for $R=2$, visual inspection of Fig.~\ref{fig:res_G_A-V}[top] confirms
that only the 4D-UWR-SENSE algorithm allows us to retrieve
significant bilateral activations in the parietal cortices~(see axial MIP slices)
in addition to larger cluster extent and a gain in significance level
for the stable clusters across the different reconstructors.
Similar conclusions can be drawn when looking at Fig.~\ref{fig:res_G_A-V}[bottom]
for $R=4$.
Complementary results are available in Table~\ref{tab:StatRes2allGrA-V}
for $R=2$ and $R=4$ and numerically confirms this visual comparison:
\begin{itemize}
\item Whatever the reconstruction method in use, the statistical performance
is much more significant using $R=2$, especially at the cluster level since the
cluster extent decreases by one order of magnitude.
\item Voxel and cluster-level results are enhanced using the 4D-UWR-SENSE approach
instead of the \texttt{mSENSE}~reconstruction or its early UWR-SENSE version.
\end{itemize}
\begin{table}[!ht]
\centering
\caption{Significant statistical results at the group-level for the aC-aS\XS contrast (corrected for multiple comparisons at $p=0.05$).
Images were reconstructed using the \texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE algorithms for $R=2$ and $R=4$.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\cline{3-7}
\cline{3-7}
\multicolumn{2}{c}{}&\multicolumn{2}{|c|}{cluster-level}&\multicolumn{3}{|c|}{voxel-level}\\
\cline{3-7}
\multicolumn{2}{c|}{}&p-value&Size&p-value&T-score& Position\\
\hline
\multirow{9}{*}{$R=2$}&\multirow{3}{*}{\texttt{mSENSE}} &$< 10^{-3}$ & 361 & 0.014&7.68&-6 -22 45\\
\cline{3-7}
& &$< 10^{-3}$ &331 & 0.014&8.23&-40 -38 42\\
\cline{3-7}
& &$< 10^{-3}$ &70 & 0.014&7.84&-44 6 27\\
\cline{2-7}
\cline{2-7}
&\multirow{3}{*}{UWR-SENSE} &$< 10^{-3}$ & 361&0.014& 7.68&-6 22 45\\
\cline{3-7}
& &$< 10^{-3}$&331&0.014& 7.68 &-44 -38 42\\
\cline{3-7}
& &$< 10^{-3}$ & 70&0.014& 7.84 &-44 6 27\\
\cline{2-7}
\cline{2-7}
&\multirow{3}{*}{4D-UWR-SENSE} &$< 10^{-3}$& \textbf{441} & $< 10^{-3}$&\textbf{9.45}&-32 -50 45\\
\cline{3-7}
& &$< 10^{-3}$ & 338 &$< 10^{-3}$&9.37&-6 12 45 \\
\cline{3-7}
& &$< 10^{-3}$ & 152 & 0.010&7.19&30 -64 48 \\
\hline
\hline
\multirow{6}{*}{$R=4$}&\multicolumn{1}{|c|}{\texttt{mSENSE}} &0.003& 14& 0.737&5.13&-38 -42 51\\
\cline{2-7}
\cline{2-7}
&\multirow{2}{*}{UWR-SENSE} &$< 10^{-3}$ & \textbf{41} & 0.274&5.78&-50 -38 -48 \\
\cline{3-7}
& &$< 10^{-3}$ & 32 & 0.274&5.91&2 12 54 \\
\cline{2-7}
\cline{2-7}
&\multirow{3}{*}{4D-UWR-SENSE} &$< 10^{-3}$& 37 & 0.268&\textbf{6.46}&-40 -40 54\\
\cline{3-7}
& &$< 10^{-3}$& 25& 0.268&6.37 &-38 -42 36\\\cline{3-7}
& &$< 10^{-3}$ & 18 & 0.273 & 5 & -42 8 36 \\
\hline
\end{tabular}
\label{tab:StatRes2allGrA-V}
\end{table}
Fig.~\ref{fig:res_G_Lc-Rc} reports similar group-level MIP results for $R=2$ and $R=4$
concerning the \texttt{Lc-Rc}\XS contrast. It is shown that whatever the acceleration factor $R$ in use,
our pipeline enables to detect a much more spatially extended activation area
in the motor cortex. This visual inspection is quantitatively confirmed
in Table~\ref{tab:StatRes2allGrLc-Rc} when comparing
the detected clusters using our 4D-UWR-SENSE approach with those found by
\texttt{mSENSE}, again irrespective of $R$. Finally, the 4D-UWR-SENSE algorithm outperforms
the UWR-SENSE one, which corroborates the benefits of the proposed spatio-temporal
regularization scheme.
\begin{figure}[!ht]
\centering
\begin{tabular}{c c c c}
&\texttt{mSENSE}&UWR-SENSE&4D-UWR-SENSE\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=2$}&\hspace*{-0.3cm}
\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R2_LC.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R2_LC.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R2_LC.eps}\\
\hspace*{-0.4cm}\raisebox{2cm}{$R=4$}&\hspace*{-0.3cm}
\includegraphics[width=3.7cm, height=3.7cm]{mSENSE_R4_LC.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{2D_R4_LC.eps}&
\hspace*{-0.3cm}\includegraphics[width=3.7cm, height=3.7cm]{4D_R4_LC.eps}\\
\end{tabular}
\caption{Group-level student-$t$ maps for the \texttt{Lc-Rc}\XS contrast where data have been reconstructed using the \texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE for $R=2$ and $R=4$.
Neurological convention. Red arrows indicate the global maximum activation peak.\label{fig:res_G_Lc-Rc}}
\end{figure}
\begin{table}[!ht]
\centering
\caption{Significant statistical results at the group-level for the \texttt{Lc-Rc}\XS contrast (corrected for multiple comparisons at $p=0.05$).
Images were reconstructed using the \texttt{mSENSE}, UWR-SENSE and 4D-UWR-SENSE algorithms for $R=2$ and $R=4$.}
\begin{tabular}{|c|c|c|c|c|c|c|}
\cline{3-7}
\cline{3-7}
\multicolumn{2}{c}{}&\multicolumn{2}{|c|}{cluster-level}&\multicolumn{3}{|c|}{voxel-level}\\
\cline{3-7}
\multicolumn{2}{c|}{}&p-value&Size&p-value&T-score& Position\\
\hline
\multirow{7}{*}{$R=2$}&
\multirow{2}{*}{\texttt{mSENSE}} &$< 10^{-3}$ & 354 &$< 10^{-3}$&9.48&38 -22 54\\
\cline{3-7}
& &0.001 &44 & 0.665&6.09&-4 -68 -24\\
\cline{2-7}
&\multirow{2}{*}{UWR-SENSE} &$< 10^{-3}$ &350& 0.005& 9.83&36 -22 57\\
\cline{3-7}
& &$< 10^{-3}$ & 35&0.286& 7.02&4 -12 51\\
\cline{2-7}
&\multirow{3}{*}{4D-UWR-SENSE} &$< 10^{-3}$& \textbf{377} & 0.001&\textbf{11.34}&36 -22 57\\
\cline{3-7}
& &$< 10^{-3}$ & 53& $< 10^{-3}$&7.50&8 -14 51 \\
\cline{3-7}
&\multicolumn{1}{|c|}{} &$< 10^{-3}$ & 47& $< 10^{-3}$&7.24&-18 -54 -18 \\
\hline
\hline
\multirow{3}{*}{$R=4$}&\multicolumn{1}{|c|}{\texttt{mSENSE}} & $< 10^{-3}$& 38& 0.990&5.97&32 -20 45\\
\cline{2-7}
&\multicolumn{1}{|c|}{UWR-SENSE} &$< 10^{-3}$ &163 &0.128&7.51&46 -18 60\\
\cline{2-7}
&\multicolumn{1}{|c|}{4D-UWR-SENSE}&$< 10^{-3}$ & \textbf{180}& 0.111& \textbf{7.61}&46 -18 60\\
\hline
\end{tabular}
\label{tab:StatRes2allGrLc-Rc}
\end{table}
\section{Discussion and conclusion}\label{sec:colclusion}
The contribution of the present paper was twofold. First, we proposed a novel reconstruction
method that relies on 3D wavelet transform and accounts for temporal dependencies in
successive fMRI volumes. Second, our particular interest was to demonstrate that when artifacts are
superimposed to brain activation, this directly impacts subsequent brain activity detection.
In this context, we showed that the choice of the parallel imaging reconstruction algorithm impacts
the statistical sensitivity in fMRI at the subject and group-levels and may enable whole brain
neuroscience studies at high spatial resolution.
Practically speaking, we showed that whole brain acquisition can be routinely used at a
spatial in-plane resolution of $2 \times 2 \mathrm{mm}^2$ in a short and constant repetition
time~($\text{TR} = 2.4$s) provided that a reliable pMRI reconstruction pipeline
is chosen. In this paper, we demonstrated that our 4D-UWR-SENSE reconstruction
algorithm meets this desired property.
To draw this conclusion, our comparison took place at the statistical analysis
level and relied on quantitative criteria~(voxel- and cluster-level corrected p-values,
$t$-scores, peak positions) both at the subject and group levels.
In particular, we showed that our 4D-UWR-SENSE contribution outperforms both its UWR-SENSE ancestor~\cite{Chaari_MEDIA_2011} and the
Siemens \texttt{mSENSE}~reconstruction in terms of statistical significance and robustness.
This emphasized the benefits of combining temporal and 3D regularizations in the
wavelet domain. Interestingly, we exhibited a most significant gain
in the more degraded situation~($R=4$) due to the positive impact of regularization.
The strength of our conclusions lies in the reasonable size of our datasets since
the same cohort participated to parallel imaging acquisitions using two
different acceleration factors~($R=2$ and $R=4$).
At this spatio-temporal compromise~($2\times2\times3$mm$^3$ and $\text{TR}=2.4$~s), we also illustrated the
impact of increasing the acceleration factor~(passing from $R=2$ to $R=4$)
on the statistical sensitivity at the subject and group levels for a given reconstruction
algorithm. We performed this comparison to anticipate what could be the statistical performance
for detecting evoked brain activity on data requiring this acceleration level,
such as high spatial resolution EPI images~(eg $1.5 \times 1.5\mathrm{mm}^2$ as
in-plane resolution) acquired at the same short $\text{TR}$.
Our conclusions were balanced depending on the contrast of
interest: when looking at the aC-aS\XS contrast, which involves in the fronto-parietal circuit,
it turned out that $R=4$ was not reliable enough to recover group-level significant activity at 3~Tesla:
the SNR loss was too important and should be compensated by an increase of the static
magnetic field~(eg passing from 3 to 7~Tesla). However, the situation is less dramatic
or even acceptable for the \texttt{Lc-Rc}\XS motor contrast, which elicits activation in motor regions:
Our results brought evidence that the 4D-UWR-SENSE approach enabled
the use of $R=4$ for this contrast.
To summarize, the compromise between the acceleration factor
and spatial in-plane resolution should be selected with care depending on the regions
involved in the fMRI paradigm. As a consequence, high resolution fMRI studies can be
conducted using high speed acquisition~(short $\text{TR}$ and large $R$ value)
provided that the expected BOLD effect is strong, as experienced in
primary motor, visual and auditory cortices. Of course, the use of an optimized
reconstruction method such as the one proposed is a pre-requisite to shift
this compromise towards larger $R$ values and higher spatial resolution and could
be optimally combined with ultra high magnetic fields.
A direct extension of the present work, which is actually in progress, consists in studying the impact of tight
frames instead of wavelet basis to define shift-invariant transformation~\cite{Chaari_MEDIA_2011}.
However, unsupervised reconstruction becomes more challenging in this framework since the
estimation of hyper-parameters becomes cumbersome~(see~\cite{Chaari_TSP_2010} for details).
Ongoing work will concern the combination of the present contribution with the joint detection estimation
approach of evoked activity~\cite{Makni08,Vincent10} to go beyond the GLM framework and measure
how the pMRI reconstruction algorithm also impacts HRF estimation.
Another extension would concern the combination of our wavelet-regularized reconstruction with the
WSPM approach~\cite{VanDeVille_2007} in which statistical analysis is directly performed in the
wavelet transform domain.
|
\section{Introduction}
Recent evidence from quasar absorption spectra suggests that there may be a spatial gradient in values of the fine-structure constant, $\alpha=e^2/\hbar c$, across cosmological distance scales~\cite{webb10arxiv}. The data come from some $\sim 300$ different absorbers using two different telescopes, Keck and the Very Large Telescope, which together provide coverage of the whole sky. All existing astronomical data measuring $\alpha$-variation are consistent with the existence of a constant spatial gradient in $\alpha$ (dipole model)~\cite{berengut11jpcs}.
The existence of a spatial dipole in $\alpha$ could be confirmed using terrestrial clocks~\cite{berengut10arxiv0}. The motion of the Sun relative to the measured dipole (that is, towards a region of the Universe with larger $\alpha$) implies a time-variation of around $\dot\alpha/\alpha \sim 10^{-18} - 10^{-19}\,\textrm{yr}^{-1}$. The annual motion of the Earth modulates the signal, giving an additional shift $\delta\alpha/\alpha = 1.4\E{-20}\cos\omega t$, where $\omega$ refers to the frequency of the annual orbit. To measure this shift, one must compare two clocks with different sensitivities to $\alpha$ over the course of several years~\cite{dzuba99prl}. The best current limit on terrestrial time-variation, $\dot\alpha/\alpha = (-1.6\pm2.3)\E{-17}\textrm{yr}^{-1}$, comes from comparison of Hg$^+$ and Al$^+$ optical clocks over the course of a year~\cite{rosenband08sci}. Here the Hg$^+$ transition is strongly dependent on $\alpha$, while the Al$^+$ clock is relatively insensitive~\cite{dzuba99prl,dzuba99pra}.
To confirm the dipole model of $\alpha$-variation in the laboratory requires finding systems where the sensitivity to $\alpha$ is enhanced. The sensitivity is indicated by $q$-coefficients, usually defined by
\begin{equation}
\label{eq:q_def}
\omega = \omega_0 + q x
\end{equation}
where $x = (\alpha/\alpha_0)^2 - 1 \approx 2(\alpha - \alpha_0)/\alpha_0$ is the change in $\alpha$ over time from its current value $\alpha_0$, which leads to a change in the frequency $\omega$. Here the atomic unit of energy, which is cancelled out in any measured ratio of frequencies, is assumed to be constant. The dependence of $\omega$ on $\alpha$ is due to relativistic corrections.
Potential clocks that have large $q$-values include optical transitions in Yb$^+$~\cite{porsev09pra} and Th$^{3+}$~\cite{flambaum09pra}, as well as the thorium ``nuclear clock''~\cite{peik03epl}, which would make use of the 7.5 eV nuclear transition in $^{229}$Th to produce a clock with a $q$-value many orders-of-magnitude larger than the optical Hg$^+$ clock~\cite{flambaum06prl,berengut09prl}. Recently, we showed that trapped, highly-charged ions could provide optical atomic clocks with much larger $q$-values than would be possible using near-neutral ions~\cite{berengut10prl}. Clocks using suggested transitions in Sm$^{14+}$, for example, would benefit from several separate contributions to $\alpha$-sensitivity: high nuclear charge $Z$, high ionization stage, and significant differences in the configuration composition of the states involved.
In this paper\ we demonstrate that using holes can dramatically increase sensitivity to $\alpha$-variation, while retaining all of the other sources of enhancement. We show that it is possible to find hole transitions in highly ionized atomic systems that are within the range of optical lasers, and discuss the sources of the large $q$ values. Two promising systems are identified along the ionization sequences of iridium and tungsten. The Ir$^{16+}$ and Ir$^{17+}$ ions are shown to have $q$-values much larger than any that have previously been seen in atomic systems.
\section{Theory}
In Refs.~\cite{berengut10prl,dzuba99pra} it was shown that for a single electron above closed shells, the relativistic shift is
\begin{equation}
\label{eq:q_external_electron}
q \approx -I_n \frac{(Z\alpha)^2}{\nu(j+\nicefrac{1}{2})}
\end{equation}
where $I_n = Z_a^2/2\nu^2$ is the ionisation energy of the orbital (atomic units $\hbar = e = m_e = 1$). Here $Z_a$ is the effective charge that the external electron ``sees'' and $\nu$ is the effective principal quantum number. It was shown in \cite{berengut10prl} that, for a valence orbital in an ion with large enough $Z$, $q \sim I_n$ as $\nu \rightarrow n$, the principal quantum number. The ratio $q_\textrm{norm} = -q/(Z^2\alpha^2 I_n)$ was computed for the $5s$ valence orbital along the Ag isoelectronic sequence; we reproduce the results (Fig.~1. of~\cite{berengut10prl}) as the filled circles in ~\Fig{fig:q5s_norm}, below.
Consider how \eref{eq:q_external_electron} might be modified for a hole. For our purposes, the hole state has the same quantum numbers as an electron in the closed shells. The hole state will have an enhanced sensitivity to $\alpha$-variation since the ionization potential is larger when there are more electrons in the outer shell, and is maximal for electrons in a closed shell. Indeed, $Z_a$ for an electron in closed shells would not be well described by $Z_a = Z_\textrm{ion}+1$ as it was previously; rather such an electron would spend approximately half its time closer to the nucleus than the other electrons in the same shell, and this leads to a larger relativistic shift. In such a case $q \sim I_n^{3/2}$ will better describe the system.
The open squares in \Fig{fig:q5s_norm} show the ratio $q_\textrm{norm}$ for the $5s$ orbital along the sequence $5s^2 4f^k$ with constant $Z_\textrm{ion}$, computed in the Dirac-Fock approximation.
The sequence starts at $5s^2$ ($k=0$) at $Z=60$, and ends at $5s^2 4f^{14}$ at $Z=74$. For all ions in the sequence $Z_\textrm{ion}=12$. Along this sequence the $4f$ orbital energies lie above the $5s$ orbital energies by less than a few percent -- in this sense the orbitals are in the same ``shell''.
\begin{figure}[tb]
\includegraphics[width=0.44\textwidth]{q5s_norm}
\caption{\label{fig:q5s_norm} Ratio $q_\textrm{norm} = -q/(Z^2\alpha^2 I_n)$ for the $5s$ orbital, calculated using the Dirac-Fock theory. Filled circles: $q_\textrm{norm}$ for the $5s$ valence orbital in the Ag isoelectronic sequence. Open squares: $q_\textrm{norm}$ for a $5s$ orbital from the configuration $5s^2 4f^k$ where the ionization degree is always $Z_{ion} = 12$.}
\end{figure}
\Fig{fig:q5s_norm} shows that there is indeed an extra contribution to the relativistic shift when the external orbital has many electrons in it: $q$ is increasing much faster than $I_n$. In the Hartree-Fock approximation for filled subshells, the ionization energy (to create a hole) is equal to the orbital energy. We see that a single hole in an otherwise filled external shell will have the highest sensitivity to variation of $\alpha$ possible given a particular $Z$ and $Z_a$. It is for this reason that we consider hole transitions as candidates for optical clocks.
To exploit the enhanced sensitivity to $\alpha$-variation afforded by hole transitions, we must find examples that have high $Z$, are highly ionized, and lie within the optical range. This is not trivial, since the difference in ionization potential between different shells increases as $\sim Z_a^2$, which rapidly takes any transition between the shells outside the range of lasers as $Z_a$ increases. To combat this tendency we must find particular examples near crossing points between the shells. The Coulomb crossing happens because in neutral atoms shells with larger angular momentum, $l$, can have energies significantly above shells with smaller momentum but larger principal quantum number, $n$ (for example, in neutral thorium the $5f$ orbital is above $7s$). On the other hand, in the hydrogenlike limit all shells with the same $n$ are nearly degenerate, regardless of angular momentum (e.g.~$E_{5f} = E_{5s}$). In this limit, a higher principal quantum number is necessarily associated with a larger energy. Therefore as $Z$ increases between the two limits there must be a crossing point where the two shells with different $n$ and $l$ have similar energies. Near such a point it may be possible to find hole-transitions within the range of lasers.
\section{Iridium}
In this paper\ we identify two separate crossing points occurring in high-$Z$ ions. The first happens between the $4f^{14}$ and $5s^2$ shells in Sm-like ions (number of electrons $N=62$) with $Z\gtrsim 70$. In \Fig{fig:E_N62} we show the Dirac-Fock energies of these closed, outermost shells as a function of $Z$. The crossing is seen to occur around $Z=77$ (iridium). In the Dirac-Fock calculation a single hole has energy given by the ionization potential, $-E$. While the ionization potentials of both shells are around $380$~eV, the transition energy between configurations of Ir$^{16+}$ having the hole in the $4f$ level ($5s^2 4f^{13}$) and the hole in the $5s$ level ($5s\,4f^{14}$) will be very much less than this. Configuration interaction calculations are presented in \Tref{tab:Ir16+}. Note that for $Z$ below $70$ the $5p$ shell can begin to fill before the $4f$ shell. This is not necessarily a bad thing: for some $Z$ and some $N < 62$ there will be crossing between partly-filled $5p$ and $4f$ shells that may be exploited to find optical transitions, however we have not considered these many-electron cases in this work because no additional advantage is gained from more complex cases.
\begin{figure}[tb]
\includegraphics[width=0.47\textwidth]{E_N62}
\caption{\label{fig:E_N62} Dirac-Fock energies of $4f_{7/2}$ (solid) and $5s$ (dashed) orbitals in the Sm isoelectronic sequence, $N=62$, with closed external shells $5s^2 4f^{14}$.}
\end{figure}
\begin{table}[tb]
\caption{Energy levels and sensitivity coefficients ($q$) for Ir$^{16+}$ (\ensuremath{\textrm{cm}^{-1}}).}
\label{tab:Ir16+}
\begin{ruledtabular}
\begin{tabular}{lcrr}
\multicolumn{1}{c}{Configuration} & $J$ &
\multicolumn{1}{c}{Energy} &
\multicolumn{1}{c}{$q$} \\
\hline
$4f^{13}5s^2\ ^2\!F^o$ & 7/2 & 0 & 0 \\
& 5/2 & 25\,898 & 23\,652 \\
$4f^{14}5s\ ^2\!S$ & 1/2 & 37\,460 & 367\,315 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\Tref{tab:Ir16+} and \Tref{tab:Ir17+} show configuration interaction (CI) calculations for low-lying levels in Ir$^{16+}$ and Ir$^{17+}$: the one-hole and two-hole cases, respectively. Details of the CI calculations have been presented elsewhere (see, e.g.~\cite{dzuba99pra,berengut04praB}). The $q$-values are calculated by repeating the entire calculation for $x=-0.01$ and $0.01$ and taking the gradient of the frequency, i.e.:
\[
q = \frac{\omega_{\delta x} - \omega_{-\delta x}}{2\,\delta x} \,,
\]
with $\delta x = 0.01$.
The resulting $q$ values are the largest ever seen in an atomic system.
Energy levels are accurate to 6000\,\ensuremath{\textrm{cm}^{-1}}, mostly due to two-electron excitations in the CI calculation not being fully saturated. However, the fine-structure splitting within terms is much more accurate than this, and the $q$-values are very stable and certainly correct to within 10\%.
\begin{table}
\caption{Energy levels and sensitivity coefficients ($q$) for Ir$^{17+}$ (\ensuremath{\textrm{cm}^{-1}}). Uncertainties between terms are at the level $\sim 6000\,\ensuremath{\textrm{cm}^{-1}}$, and in particular it is possible that the ground state is actually $4f^{14}\ ^1\!S_0$. However, since this level is metastable, it does not qualitatively affect our analysis of this system.}
\label{tab:Ir17+}
\begin{ruledtabular}
\begin{tabular}{lcrr}
\multicolumn{1}{c}{Configuration} & $J$ &
\multicolumn{1}{c}{Energy} &
\multicolumn{1}{c}{$q$} \\
\hline
\multicolumn{4}{c}{Odd states} \\
$4f^{13}5s\ ^3\!F^o$ & 4 & 0 & 0 \\
$4f^{13}5s\ ^3\!F^o$ & 3 & 4\,838 & 2\,065 \\
$4f^{13}5s\ ^3\!F^o$ & 2 & 26\,272 & 24\,183 \\
$4f^{13}5s\ ^1\!F^o$ & 3 & 31\,492 & 25\,052 \\
\multicolumn{4}{c}{Even states} \\
$4f^{14}\ ^1\!S$ & 0 & 5\,055 & 367\,161 \\
$4f^{12}5s^2\ ^3\!H$ & 6 & 35\,285 & -385\,367 \\
$4f^{12}5s^2\ ^3\!F$ & 4 & 45\,214 & -387\,086 \\
$4f^{12}5s^2\ ^3\!H$ & 5 & 59\,727 & -362\,127 \\
$4f^{12}5s^2\ ^3\!F$ & 2 & 68\,538 & -378\,554 \\
$4f^{12}5s^2\ ^1\!G$ & 4 & 68\,885 & -360\,678 \\
$4f^{12}5s^2\ ^3\!F$ & 3 & 71\,917 & -362\,313 \\
$4f^{12}5s^2\ ^3\!H$ & 4 & 92\,224 & -339\,253 \\
$4f^{12}5s^2\ ^1\!D$ & 2 & 98\,067 & -363\,983 \\
$4f^{12}5s^2\ ^1\!J$ & 6 & 110\,065 & -364\,732 \\
$4f^{12}5s^2\ ^3\!P$ & 0 & 110\,717 & -372\,570 \\
$4f^{12}5s^2\ ^3\!P$ & 1 & 116\,372 & -362\,937 \\
\end{tabular}
\end{ruledtabular}
\end{table}
Let us take two particular examples from Ir$^{17+}$, the two-hole case. Trapping and cooling of this ion may be facilitated by any of the E1 transitions from the ground state $4f^{13}5s\ ^3\!F_4^o$, such as $4f^{12}5s^2\ ^3\!F_4$ at $\omega \sim 45\,000\,\ensuremath{\textrm{cm}^{-1}}$. While this transition itself has an extremely large $q$-value, and could therefore in principle be used to test $\alpha$-variation, E1 transitions tend to be too broad for precise measurements. On the other hand the transition to $4f^{12}5s^2\ ^3\!H_6$ with frequency $\sim 35\,000\,\ensuremath{\textrm{cm}^{-1}}$ could be induced via hyperfine mixing with its fine-structure partner $^3H_5$, and this transition would have a strongly reduced linewidth (it may also go as E3/M2). The most abundant isotope of iridium is $^{193}$Ir, which is stable and has nuclear spin $I = 3/2$. Therefore the hyperfine-induced transitions are always available.
The $4f^{14}\ ^1\!S_0$ line is interesting because the $\alpha$-sensitivity acts in the opposite direction ($q$ is large and positive). However, when trying to use this line one expects to encounter the problem of too little line strength. The best option here is to use the transition from the very metastable $4f^{14}\ ^1\!S_0$ to $4f^{13}5s\ ^3\!F_2^o$ ($\omega \approx 21\,000\,\ensuremath{\textrm{cm}^{-1}}$). This is an M2/E3 transition, and at first seems too slow for existing lasers. However, due to the hyperfine mixing with $J=1$ states such as the $116\,000\,\ensuremath{\textrm{cm}^{-1}}$ $4f^{12}5s^2\ ^3\!P_1$ level, this transition may also proceed via hyperfine-induced E1 transition. Odd-parity $J=1$ states also exist; hyperfine mixing with these states would be dominated by the configuration $4f^{12}5s5p$ at energies $\sim 500\,000\,\ensuremath{\textrm{cm}^{-1}} \sim 60$\,eV, which is still far below the continuum of such a highly-ionized system.
The best current limit on terrestrial time-variation of $\alpha$ comes from an experiment comparing Hg$^+$ and Al$^+$ clocks~\cite{rosenband08sci}, with $\Delta q = 57\,000\,\ensuremath{\textrm{cm}^{-1}}$~\cite{dzuba99pra}. By contrast, an experiment that compared the two transitions we have examined ($4f^{14}\ ^1\!S_0$ to $4f^{13}5s\ ^3\!F_2^o$ and $4f^{13}5s\ ^3\!F_4^o$ to $4f^{12}5s^2\ ^3\!H_6$) would have a total sensitivity to $\alpha$-variation of $\Delta q \approx 730\,000\,\ensuremath{\textrm{cm}^{-1}}$. The largest $q$ ever found in a single atomic transition occurs in $4f^{14}\ ^1\!S_0$ to $4f^{12}5s^2\ ^3\!F_2$, however in this case the upper level has an E1-decay mechanism and may be too broad.
\section{Tungsten}
The second crossing point we consider happens when the filled $4f^{14}$ and $5p^6$ shells cross in Er-like ions ($N=68$). Because of the relatively large splitting between the $5p_{3/2}$ and $5p_{1/2}$ orbitals (compared to the much smaller fine-structure splitting of the $4f$ orbitals), this crossing point is more spread out, somewhere between $Z=73$ and~75 (see \Fig{fig:E_N68}). We focus on $Z=74$ because tungsten is a well-studied element due to its use in plasma diagnostics (see, e.g.~\cite{clementson10jpb}), and hence may be of higher interest to experimenters. CI calculations for the one-hole and two-hole systems, W$^{7+}$ and W$^{8+}$ respectively, are presented in Tables~\ref{tab:W7+} and~\ref{tab:W8+}.
\begin{figure}[tb]
\includegraphics[width=0.47\textwidth]{E_N68}
\caption{\label{fig:E_N68} Dirac-Fock energies of $4f$ (solid), $5s$ (dashed), and $5p$ (dot-dashed) orbitals in the Er isoelectronic sequence, $N=68$, with closed external shells $5s^2 5p^6 4f^{14}$. The orbital fine-structure splitting is indicated, with closed circles for $4f_{7/2}$ and $5p_{3/2}$ and open circles for the $4f_{5/2}$ and $5p_{1/2}$ orbitals.}
\end{figure}
\begin{table}[tb]
\caption{Energy levels and sensitivity coefficients ($q$) relative to the ground state for W$^{7+}$ (\ensuremath{\textrm{cm}^{-1}}).}
\label{tab:W7+}
\begin{ruledtabular}
\begin{tabular}{lcrbr}
\multicolumn{1}{c}{Configuration} & $J$ &
\multicolumn{2}{c}{Energy} &
\multicolumn{1}{c}{$q$} \\
& & \multicolumn{1}{c}{This work} & \multicolumn{1}{c}{\cite{kramida09adndt}} & \\
\hline
$4f^{13}5p^6\ ^2\!F^o$ & 7/2 & 0 & 0 & 0 \\
& 5/2 & 18\,199 & 17\,440 & 16\,462 \\
$4f^{14}5p^5\ ^2\!P^o$ & 3/2 & 4\,351 & 800 (700) & 87\,544 \\
& 1/2 & 93\,908 & 87\,900 (700) & 200\,269 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{table}[tb]
\caption{Energy levels and sensitivity coefficients ($q$) for W$^{8+}$ (\ensuremath{\textrm{cm}^{-1}}). Energies between terms are uncertain at the level $\sim 6000\,\ensuremath{\textrm{cm}^{-1}}$.}
\label{tab:W8+}
\begin{ruledtabular}
\begin{tabular}{lcrr}
\multicolumn{1}{c}{Configuration} & $J$ &
\multicolumn{1}{c}{Energy} &
\multicolumn{1}{c}{$q$} \\
\hline
$4f^{14}5p^4\ ^3P$ & 2 & 0 & 0 \\
$4f^{13}5p^5\ ^3F$ & 4 & 6\,075 & -81\,564 \\
$4f^{13}5p^5\ ^3G$ & 3 & 6\,357 & -81\,480 \\
$4f^{13}5p^5\ ^3G$ & 5 & 11\,122 & -82\,880 \\
$4f^{13}5p^5\ ^3F$ & 3 & 21\,905 & -66\,489 \\
$4f^{13}5p^5\ ^3D$ & 2 & 23\,276 & -66\,985 \\
$4f^{13}5p^5\ ^3F$ & 2 & 28\,112 & -66\,124 \\
$4f^{14}5p^4\ ^1S$ & 0 & 29\,810 & 12\,735 \\
$4f^{13}5p^5\ ^3G$ & 4 & 34\,884 & -65\,896 \\
$4f^{13}5p^5\ ^3D$ & 1 & 36\,497 & -65\,946 \\
$4f^{12}5p^6\ ^3H$ & 6 & 56\,416 & -179\,990 \\
$4f^{12}5p^6\ ^3F$ & 4 & 65\,008 & -181\,419 \\
$4f^{12}5p^6\ ^3H$ & 5 & 73\,188 & -164\,139 \\
$4f^{12}5p^6\ ^1G$ & 4 & 80\,551 & -162\,573 \\
$4f^{12}5p^6\ ^3F$ & 2 & 82\,424 & -177\,137 \\
$4f^{12}5p^6\ ^3F$ & 3 & 83\,315 & -164\,072 \\
\end{tabular}
\end{ruledtabular}
\end{table}
Previous studies extrapolating from the measured W$^{6+}$ spectrum~\cite{sugar75pra} have noted that the $4f$ and $5p$ orbitals compete for the ground state in W$^{7+}$~\cite{kramida09adndt}. We present these calculations in \Tref{tab:W7+} but repeat the warnings of~\cite{kramida09adndt} that the accuracy may be low. The fine-structure splittings are more accurate.
Both W$^{7+}$ and W$^{8+}$ have several M1 and E2 transitions that may be suitably narrow to be of use in studies of $\alpha$-variation. For example, the E2 transition from the ground state of W$^{8+}$, $4f^{14}5p^4\ ^3P_2$, to $4f^{12}5p^6\ ^3F_4$ has energy $\omega \sim 65\,000\,\ensuremath{\textrm{cm}^{-1}}$ and $q = -181\,000\,\ensuremath{\textrm{cm}^{-1}}$. This transition may be compared to the $4f^{14}5p^4\ ^1S_0$ transition, or to any line with a positive or small $q$ value in another ion, to give a sensitivity to $\alpha$-variation which would still be the largest ever utilised. By comparing instead to the $4f^{13}5p^6\ ^2\!F^o_{5/2}$ to $4f^{14}5p^5\ ^2\!P^o_{1/2}$ E2 transition in W$^{7+}$, one could achieve $\Delta q \approx 365\,000\,\ensuremath{\textrm{cm}^{-1}}$.
\section{Conclusion}
We have shown that level-crossings between filled shells provide an opportunity to find hole-transitions within laser range in highly-charged ions. These hole transitions can have hugely enhanced sensitivity to $\alpha$-variation when compared with transitions used today. As atomic spectroscopy in electron-beam ion traps continues to improve (see, e.g.~\cite{draganic03prl,crespo08cjp,hobein11prl} and review~\cite{beiersdorfer09pscr}) we expect that hole-transitions in multiply-charged ions may become good systems for searching for time-variation of $\alpha$. We have found candidates in two different systems that may be of interest, and suggest that others may exist along other isoelectronic sequences. Level crossings also allow us to identify optical-range E1 transitions such as those we have found in Ir$^{17+}$ that may be of importance in laser-trapping and cooling of these ions.
\acknowledgments
This work was supported by the Australian Research Council. Supercomputer time was provided by an award under the Merit Allocation Scheme on
the NCI National Facility at the Australian National University.
|
\subsection*{Minimization of the scalar potential}
The explicit equations for the minimum of the scalar potential considered above are shown in what follows. In a first approximation we neglect all the terms suppressed by $\epsilon_{u,d,\theta}$.
\begin{itemize}
\item
The equation associated with $y_u$ is then given by:
\begin{equation}
\dfrac{\partial V_\Sigma}{\partial y_u}=2y_u\Lambda_f^4\left[-\frac{\mu_u^2}{\Lambda_f^2}+2\lambda_u\left(y_u^2+y_c^2\right)+\tilde{\lambda}_u y_c^2-2\lambda_{udud}F_u\right]=0\,.
\label{eqminyu}
\end{equation}
The term $F_u$ is a function of parameters that will not enter in the determination of $y_u$ and vanishes in the limit of massless first family and no mixing.
The physical choice in eq. (\ref{eqminyu}) is to cancel the first factor taking $y_u=0$, as the cancellation of the other factor would lead to $y_c=0$. This is stable provided that $\tilde{\lambda}_u>0$. In a similar way, $y_d=0$ is a solution to the equation $\partial V_ \Sigma/\partial y_d=0$ .
\item
When deriving with respect to the angle $\theta$ we find
\begin{equation}
\begin{split}
\dfrac{\partial V_\Sigma}{\partial \theta}=&2\sin{2\theta}\,\lambda_{udud}\left(y_c^2-y_u^2\right)\left(y_s^2-y_d^2\right)\times\\
&\times\left[\left(y_c^2-y_u^2\right)\left(y_s^2-y_d^2\right)\sin^2\theta+\left(y_c^2y_d^2+y_s^2y_u^2\right)\right]=0\,.
\end{split}
\label{eqanglec}
\end{equation}
Substituting the solutions to the previous minima equations considered, $y_u=0=y_d$, eq. (\ref{eqanglec}) forces $\sin{2\theta}=0$.
\item
For the heavy Yukawa couplings, once $y_u=y_d=0$ is chosen the equations take the form:
\begin{equation}
\frac{\partial V_\Sigma}{\partial y_c}=2y_c\,\Lambda_f^4\left(2\,\lambda_u\,y_c^2-2\,\lambda_{udud}\,y_c^2\,y_s^4-\frac{\mu_u^2}{\Lambda_f^2}\right)=0\,.
\end{equation}
Neglecting the trivial solution, which is unstable for positive definite coefficients, this equation yields the expression for $y_c$:
\begin{equation}
y_c^2=\frac{\mu_u^2}{2\Lambda_f^2\left(\lambda_u-\lambda_{udud}\,y_s^4\right)}\simeq\frac{\mu_u^2}{2\Lambda_f^2\lambda_u}\,,
\end{equation}
where the last equality holds when taking into account the observed value of the strange Yukawa coupling. A similar result can be found for $y_s$.
\end{itemize}
Summarizing, neglecting all terms suppressed by $\epsilon_{u,d,\theta}$, the minimum of the scalar potential is given by:
\begin{equation}
y_u=y_d=0\,,\qquad \quad
y_c=\dfrac{\mu_u}{\sqrt{2}\Lambda_f\sqrt{\lambda_u}}\,,\qquad \quad
y_s=\dfrac{\mu_d}{\sqrt{2}\Lambda_f\sqrt{\lambda_d}}\,,\qquad \quad
\sin\theta=0\,.
\end{equation}
The observed values of $y_c$ and $y_s$ Y are understood as the outcome of the hierarchy among the vevs of the flavons, $\langle\chi\rangle\sim\mu$, and the flavour scale $\Lambda_f$. Note that the parameters $\epsilon_{u,d,\theta}$ do not enter into the definition of $y_c$ and $y_s$, but control the hierarchy between the light and the heavy generations and the appearance of a non-trivial mixing angle. This solution is stable with all the coefficients in eq. (\ref{proppot}) positive and furthermore the inclusion of the corrections given by the $\epsilon$-terms will shift the minimum but will not change its stability.\\
We now discuss the changes of the solutions found above by the introduction of $\epsilon_{u,d,\theta}$.
\begin{itemize}
\item
The corrections for the first family Yukawa couplings shift their values from zero by an amount $\epsilon_{u,d}$. Explicitly, once the leading order solution found for $y_c$ and $y_s$ is inserted into eq. (\ref{eqminyu}), the dominant contributions are given by
\begin{equation}
\dfrac{\partial V_\Sigma}{\partial y_u}=\Lambda_f^4\left[2y_u\tilde{\lambda}_u y_c^2-\epsilon_u\frac{\tilde{\mu}_u^2}{\Lambda_f^2}y_c\right]=0\,,
\end{equation}
which leads to a non-vanishing Yukawa coupling for the up quark:
\begin{equation}
y_u=\epsilon_u\frac{\sqrt{\lambda_u}\,\tilde{\mu}_u}{\sqrt{2}\,\tilde{\lambda}_u\,\mu_u}\frac{\tilde\mu_u}{\Lambda_f}\,.
\end{equation}
A similar result holds also for $y_d$.
\item
When considering the equation that determines the mixing angle, several corrections are present, although the dominant one is given by
\begin{equation}
\dfrac{\partial V_\Sigma}{\partial \theta}=2\sin{\theta}\cos{\theta}y_c^2y_s^2\left[2\lambda_{udud}\left(y_c^2y_s^2\sin^2\theta\right)-\epsilon_\theta\lambda_{ud}\right]=0\,,
\end{equation}
and the corresponding non-trivial solution reads
\begin{equation}
\sin^2\theta=\epsilon_\theta\dfrac{\lambda_{ud}}{2\,\lambda_{udud}\,y_s^2\,y_c^2}\,.
\end{equation}
\end{itemize}
The minimum of the scalar potential proposed in eq. (\ref{proppot}) is then given by
\begin{eqnarray}
\begin{gathered}
y_u \simeq \epsilon_u\frac{\sqrt{\lambda_u}\,\tilde{\mu}_u}{\sqrt{2}\,\tilde{\lambda}_u\,\mu_u}\frac{\tilde\mu_u}{\Lambda_f}\,,\qquad\qquad
y_d \simeq \epsilon_d\frac{\sqrt{\lambda_d}\,\tilde{\mu}_d}{\sqrt{2}\,\tilde{\lambda}_d\,\mu_d}\frac{\tilde\mu_d}{\Lambda_f}\,,\\[4mm]
y_c \simeq \frac{\mu_u}{\sqrt2\,\Lambda_f\,\sqrt\lambda_u}\,,\qquad\qquad
y_s \simeq \frac{\mu_d}{\sqrt2\,\Lambda_f\,\sqrt\lambda_d}\,,\\[4mm]
\sin^2\theta \simeq \epsilon_\theta\dfrac{\lambda_{ud}}{2\,\lambda_{udud}\,y_s^2\,y_c^2}\,.
\end{gathered}
\label{solbifund}
\end{eqnarray}
We can now specify the value of $\epsilon_{u,d,\theta}$ in order to accommodate the observed hierarchies and mixing for the first two generations: considering the ratios $\mu/(\sqrt\lambda\Lambda_f)\approx\tilde\mu/(\sqrt{\tilde\lambda}\Lambda_f)\sim10^{-3}$,
it follows that
\begin{equation}
\epsilon_u\sim 10^{-3}\,,\qquad \qquad
\epsilon_d\sim 5\times10^{-2}\,,\qquad \qquad
\epsilon_\theta\sim 10^{-10}\,,
\end{equation}
must hold.
A comment is in order: when discussing this special illustrative scalar potential, we considered up to dimension 8 operators, while neglecting many terms otherwise allowed by the symmetry. However, even such an arbitrary choice was not sufficient to recover realistic mass hierarchies and the mixing angle, and further fine-tunings were required, including $\epsilon$ values as tiny as $10^{-10}$ to recover the Cabibbo angle. These remarks should suffice to show how unnatural is the set up when trying to fix all observables from pure $d=5$ Yukawa operators.
\subsection*{Three Family Case}
The three family case involves a wider variety of operators. This is because some of the accidental simplifications in two families no longer hold for three. The analytic treatment to find the minima becomes more complicated as well, as the number of observables increases to six quark masses and three angles (obviating the CP-odd phase). We present a graphic analysis of the scalar potential in this case. This approach assumes a positive definite potential whose minimum is just the point in which the geometrical surfaces defined by constant invariant quantities meet. When focusing on the masses in either the up or the down sector, we project the parameter space to one that has as many dimensions as families. This means that instead of the curves in the $(y_c,y_u)$ plane of figure \ref{grapview} we will consider surfaces in $(y_t,y_c,y_u)$ space.
The lowest dimension invariants that involve Yukawa eigenvalues only for the up sector correspond to:
\begin{equation}
\begin{array}{l}
\mean {A_u}=\Lambda_f^2\,\left(y_t^2+y_c^2+y_u^2\right)\,,\\[2mm]
\mean {B_u}=\Lambda_f^3\,y_t\,y_c\,y_u\,,\\[2mm]
\mean {A'_{uu}}=\mean{A_u^2-A_{uu}}=2\,\Lambda_f^4\left(y_t^2\,y_c^2+y_u^2\,y_t^2+y_c^2\,y_u^2\right)\,,
\end{array}
\label{inv3fam}
\end{equation}
where the last invariant is introduced as a linear combination of some of those in eq. (\ref{Invariant3F_Fund}). Notice that three independent invariants are necessary to fix the three different masses. We can study the intersection of the surfaces defined by giving fixed values to these operators.
\begin{figure}[h]
\begin{minipage}{0.3 \textwidth}
\begin{center}
\includegraphics[width=1.5in]{3d3fam1}
\caption{Surface of constant $A_{u}$ in $\{y_u, y_c, y_t\}$ space.}
\label{fig1}
\end{center}
\end{minipage}
\hfill \begin{minipage}{0.3\textwidth}
\begin{center}
\includegraphics[width=1.5in]{3d3fam2}
\caption{Surface of constant $A'_{uu}$ in $\{y_u, y_c, y_t\}$ space.}
\label{fig2}
\end{center}
\end{minipage}
\hfill \begin{minipage}{0.3\textwidth}
\begin{center}
\includegraphics[width=1.5in]{3d3fam3}
\caption{Surface of constant $B_{u}$ in $\{y_u, y_c, y_t\}$ space.}
\label{fig3}
\end{center}
\end{minipage}
\end{figure}
In view of these surfaces and the expressions of the invariants, the vevs of the fields shall satisfy the hierarchy:
\begin{equation}
\frac{\left\langle B_u\right\rangle}{\Lambda_f^3}\ll\frac{\left\langle A'_{uu}\right\rangle}{\Lambda_f^4}\ll\frac{\left\langle A_u\right\rangle}{\Lambda_f^2}\,.
\label{hierarU}
\end{equation}
The same analysis for the down sector leads to the following relation:
\begin{equation}
\frac{\left\langle B_d\right\rangle}{\Lambda_f^3}\sim\frac{\left\langle A'_{dd}\right\rangle}{\Lambda_f^4}\ll\frac{\left\langle A_d\right\rangle}{\Lambda_f^2}\,.
\label{hierarD}
\end{equation}
\begin{minipage}{0.6\textwidth}
The geometrical analysis allows to interpret the vevs of the invariant operators as geometric quantities, assuming the hierarchy in eqs. (\ref{hierarU}), (\ref{hierarD}); as can be seen in the figure:
\begin{enumerate}
\item $\left\langle A_i\right\rangle$ sets the radius of the sphere, therefore sets the value of the highest mass.
\item The value of $\left\langle A'_{ii}\right\rangle$ determines how close is the surface in figure \ref{fig2} to the axis. The intersection of this curve and the sphere is a circle around the axis, and the radius of such circle is related to the second highest value of mass.
\item The quantity $\left\langle B_i\right\rangle$ sets the distance of the surface shown in figure \ref{fig3} to the planes $y_t\,y_c$, $y_c\,y_u$ and $y_u\,y_t$. This surface, considered in the plane of the circle determined by the intersection of the previous surfaces, is an hyperbola so that the graphic image connects to that for the case of two families.
\end{enumerate}
\end{minipage}
\begin{minipage}{0.35\textwidth}
\begin{center}
\includegraphics[width=\textwidth]{3d3fam4}
Determination of the minimum for a positive definite potential constructed with the invariants in eq. (\ref{inv3fam}).
\end{center}
\end{minipage}\\
The requirements of eq.~(\ref{hierarU}) are not naturally obtained from a general potential, the typical ansatz
to fix the vev of the invariants being through a ``double-well'' potential of the type:
\begin{equation}
V_{f-t}=\lambda_u\left(A_u-v_{A_u}\right)^2+\frac{\gamma_u}{\Lambda_{fl}^3}\left(B_u-v_{B_u}\right)^2
+\frac{\gamma'_u}{\Lambda_{fl}^4}\left(A_{uu}'-v_{A_{uu}'}\right)^2 \,.
\end{equation}
However, for writing this kind of potential, one has to neglect many cross terms that would typically spoil the hierachy.
Again, the argument proposed here only illustrates a possible, clearly not natural, way to fix the quark masses.
The mixing angles appear in the potential through the operator $A_{ud}$:
\begin{equation}
\begin{array}{l}
V^{(4)}\supset\lambda_{ud}A_{ud}=\lambda_{ud}\left(P_0+P_{int}\right)\,,\\[3mm]
P_0=-\sum_{i<j} \left(y_{u_i}^2-y_{u_j}^2 \right) \left( y_{d_i}^2-y_{d_j}^2 \right) \sin^2 \theta_{ij}\,,\\[3mm]
\begin{split}
P_{int}=&\sum_{i<j,k}\left(y_{d_i}^2-y_{d_k}^2 \right) \left( y_{u_j}^2-y_{u_k}^2 \right) \sin^2 \theta_{ik}\sin^2 \theta_{jk}+\\
&-\left( y_d^2-y_s^2 \right)\left( y_c^2-y_t^2 \right) \sin^2\theta_{12} \sin^2\theta_{13} \sin^2\theta_{23}+ \\
&+\dfrac{1}{2}\left( y_d^2 -y_s^2\right)\left( y_c^2-y_t^2 \right)\cos\delta\, \sin2\theta_{12}\sin2\theta_{23}\sin\theta_{13}\,.
\end{split}
\end{array}
\end{equation}
Neglecting the Yukawa couplings for the first family, the equations determining the angles at the minimum of the potential are given by
\begin{equation}
\begin{array}{l}
c_{12}c_{23}s_{12}s_{23}s_{13}\sin\delta=0\,,\\[2mm]
s_{12}c_{12}\left[y_c^2+y_t^2\left(s_{23}^2-s_{13}^2c_{23}^2\right)\right]-y_t^2s_{13}s_{23}c_{23}\left(c^2_{12}-s^2_{12}\right)\cos{\delta}=0\,,\\[2mm]
s_{23}c_{23}\left[y_b^2-y_b^2s_{13}^2+y_s^2s^2_{12}\left(1+s^2_{13}\right)\right]-y_s^2s_{13}s_{12}c_{12}\left(c^2_{23}-s^2_{23}\right)\cos{\delta}=0\,,\\[2mm]
s_{13}c_{13}\left(1-s_{23}^2\right)\left(y_b^2-y_s^2s^2_{12}\right)-y_s^2s_{12}c_{12}s_{23}c_{23}c_{13}\cos{\delta}=0\,,
\end{array}
\end{equation}
where $c_{ij}$ and $s_{ij}$ stand for $\cos\theta_{ij}$ and $\sin\theta_{ij}$. The last three equations can be combined into:
\begin{eqnarray}
&&s^2_{12}c^2_{12}\left[y_c^2+y_t^2\left(s_{23}^2-s_{13}^2c_{23}^2\right)\right]^2=y_t^4s_{13}^2s_{23}^2c_{23}^2(1-2s_{12}^2)^2\cos^2{\delta}
\label{eqangle12}\,,\\
&&s_{23}^2\left(y_b^2+y_b^2s_{13}^2+y_s^2 s_{12}^2c_{13}^2\right)=s_{13}^2(y_b^2-y_s^2 s_{12}^2)
\label{eqangle23}\,,\\[1mm]
&&s_{13}^2c^2_{13}c^2_{23}\left(y_b^2-y_s^2 s^2_{12}\right)\left[y_b^2-y_s^2\left(\cos^2{\delta}s_{12}^2+\sin^2{\delta}s^4_{12}\right)\right]=0\,.
\label{eqangle13}
\end{eqnarray}
From eq. (\ref{eqangle13}) it follows that $\sin\theta_{13}=0$ is a solution. Neglecting this angle, $\sin\theta_{23}=0$ can be derived from eq. (\ref{eqangle23}). Finally, from eq. (\ref{eqangle12}), it would result $\sin\theta_{12}=0$. The other alternatives: $\cos\theta_{13}=0$ or $\cos\theta_{13}=0$ lead to unphysiscal solutions but stand as nonvanishing angle configurations and therefore a novel -if unrealistic- possibility with respect to the two family case.
\section*{C Scalar potential for the fundamental approach}
\mathversion{bold}
\section{Introduction}
After years of intense searches, all flavour processes observed in the hadronic sector, from rare decays
measurements in the kaon and pion sectors to superB--factories results, are well in agreement with the
expectations of the Standard Model of particle physics (SM). To say that all flavour processes are consistent
with the SM predictions is tantamount to state that all flavour effects observed until now are consistent with
being generated through the Yukawa couplings, which are the sole vehicles of flavour and CP violation in the SM.
Nevertheless, the origin of fermion masses and mixings remains the most unsatisfactory question in the visible
sector of nature: it involves important fine-tunings and lack of predictivity, as essentially for each mass or
mixing angle a new parameter is added by hand to the SM. It is commonly expected that an underlying dynamics will
provide a rationale for the observed patterns.
The hypothesis of {\it Minimal Flavour Violation} (MFV)~\cite{DGIS:MFV} is a humble, matter-of-fact, and highly
predictive working frame built only on: i) the assumption that, at low energies, the Yukawa couplings are the
only sources of flavour and CP violation both in the SM and in whatever may be the flavour theory beyond it,
abiding in this way to the experimental indications mentioned above; ii) the use of the flavour symmetries which
the SM exhibits in the limit of vanishing Yukawa couplings.
Indeed, the hadronic part of the SM Lagrangian, in the absence of quark Yukawa terms, exhibits a flavour symmetry given by
\begin{equation}
G_f=SU(3)_{Q_L}\times SU(3)_{U_R}\times SU(3)_{D_R}\,,
\label{3famsym}
\end{equation}
plus three extra $U(1)$ factors corresponding to the baryon number, the hypercharge and the Peccei-Quinn
symmetry\cite{PQ:U1PC}. The non-abelian subgroup $G_f$ controls the flavour structure of the Yukawa matrices,
and we focus on it for the remainder of this paper. Under $G_f$, the $SU(2)_L$ quark
doublet, $Q_L$, and the $SU(2)_L$ quark singlets, $U_R$ and $D_R$, transform as:
\begin{equation}
Q_L\sim(3,1,1)\,,\qquad\qquad
U_R\sim(1,3,1)\,,\qquad\qquad
D_R\sim(1,1,3)\;.
\label{QuarksTrans}
\end{equation}
The SM Yukawa interactions break explicitly the flavour symmetry:
\begin{equation}
\mathscr{L}_Y=\ov{Q}_LY_DD_RH+\ov{Q}_LY_UU_R\tilde{H}+\text{h.c.}
\end{equation}
The technical realization of the MFV ansatz promotes the Yukawa couplings $Y_{U,D}$ to be spurion fields
which transform under $G_f$ as
\begin{equation}
Y_U\sim(3,\ov{3},1)\;,\qquad\qquad
Y_D\sim(3,1,\ov{3})\;,
\label{spurions}
\end{equation}
recovering the invariance under $G_f$ of the full SM Lagrangian. Following the usual
MFV convention for the Yukawas, one defines
\begin{equation}
Y_D=\left(
\begin{array}{ccc}
y_d & 0 & 0 \\
0 & y_s & 0 \\
0 & 0 & y_b \\
\end{array}
\right)\;,\qquad\qquad
Y_U={\mathcal V}^\dag_{CKM}\left(
\begin{array}{ccc}
y_u & 0 & 0 \\
0 & y_c & 0 \\
0 & 0 & y_t \\
\end{array}\right)\;,
\label{SpurionsVEVs}
\end{equation}
with ${\mathcal V}_{CKM}$ being the usual quark mixing matrix, encoding three angles and one CP-odd phase.
MFV is not a model of flavour and the value of the new dynamical flavour scale $\Lambda_f$ is not fixed:
it does not determine the energy scale at which new flavour effects will show up. Nevertheless it is quite
successful in predicting precise and constrained relations between different flavour transitions, to be observed
whenever the new physics scale becomes experimentally accessible~\cite{Buras:MFVandBeyond}. The reason is that in the
MFV framework the coefficients of all SM--gauge invariant operators have a fixed flavour structure in terms of Yukawa couplings, so as to make the operator invariant under $G_f$, plus the fact that the top Yukawa coupling may dominate any coefficient in which it participates~\footnote{This is modified, though, in some MFV versions
such as two--Higgs doublet models~\cite{DGIS:MFV} with extra discrete symmetries~\cite{Branco}, or in models with strong
dynamics~\cite{Kagan}}.
MFV sheds also an interesting light on the relative size of the electroweak and the flavour scale. The origin of all
visible masses and the family structure are the two major unresolved puzzles of the SM and it is unknown whether
a relation exists between the nature and size of those two scales. While the electroweak data, and the theoretical
fine-tunings they require, suggest that new physics should appear around the TeV scale, traditional model-independent
limits on the flavour scale $\Lambda_f$ point to order(s) of magnitude larger values~\cite{INP:GenericFS}. Within MFV
both sizes could be reconciled instead around the TeV scale, due to the Yukawa suppression of the flavour-changing
operator coefficients. This holds either assuming only the SM as the
renormalizable theory~\cite{DGIS:MFV} or in beyond the SM scenarios (BSM), such as supersymmetric~\cite{LPR:MFVsusy}
or extradimensional~\cite{FPR:MFVextraD} versions of the MFV ansatz\footnote{The BSM theory may introduce more
than one distinct flavour scale:
this work sticks to a conservative and minimalist approach, focusing on the physics
related to $\Lambda_f$ as described above.}.
It is unlikely that MFV holds at all scales~\cite{Isidori2-bis}. MFV assumes a new dynamical scale $\Lambda_f$,
which points to MFV being just an accidental low-energy property of the theory. In this sense, MFV implicitly
points to a dynamical origin for the values of the Yukawa couplings. The latter may correspond to the vacuum
expectation values (vevs) of elementary or composite fields or combinations of them. In other words,
the spurions may be promoted to fields, usually called flavons. For instance, in the first formulation of MFV by
Chivukula and Georgi~\cite{CG:MFV}, the Yukawa couplings corresponded to a fermion condensate. In this work, we
further explore the dynamical character of the flavons, in a rather model--independent way.
The Yukawa interactions may be then seen as effective operators of dimension larger than four
--denominated {\it Yukawa operators} in what follows-- weighted down by powers of the large flavour
scale\footnote{For instance, a possible realization among many takes $\Lambda_f$ to be the mass
of heavy flavour mediators in some BSM theory \cite{Syms}: at energies $E<\Lambda_f$, they can be integrated out resulting
in $d>4$ operators involving the SM fields and the flavons.} $\Lambda_f$. The precise dimension $d$ of the
Yukawa operators is not determined, as illustrated in Fig.~\ref{EffectiveYukawa}. As long as the vev to be
taken by the flavon fields is smaller than $\Lambda_f$, an analysis ordered by inverse powers of this scale
is a sensible approach.
\begin{figure}[h!]
\centering
\includegraphics[height =5cm]{DibujoFlavour.pdf}
\caption{Effective Yukawa coupling.}
\label{EffectiveYukawa}
\end{figure}
The simplest case is that of a $d=5$ operator:
\begin{equation}
\mathscr{L}_Y=\ov{Q}_L\dfrac{\Sigma_d}{\Lambda_f}D_RH+\ov{Q}_L\dfrac{\Sigma_u}{\Lambda_f}U_R\tilde{H}+\text{h.c.}\;,
\label{LagrangianBi_Fund}
\end{equation}
with the scalar flavons $\Sigma_d$ and $\Sigma_u$ being dynamical fields in the bi--fundamental representation of $G_f$ (i.e.
$\Sigma_u\sim(3,\ov{3},1)$ and $\Sigma_d\sim(3,1,\ov{3}$), see eq.~(\ref{spurions})) such that\footnote{The
Goldstone bosons that would result from the spontaneous breaking of a continuous global flavour symmetry,
may be avoided for instance by gauging the symmetry. In practical realizations, this in turn tends to induce
dangerous flavour--changing neutral currents mediated by the new gauge bosons. A new promising avenues to cope
with this problem has been recently proposed in ref.~\cite{GRV:SU3gauged,Feldmann:SU5gauged}.}
\begin{equation}
Y_D \equiv \dfrac{\mean{\Sigma_d}}{\Lambda_f}\,, \qquad \qquad
Y_U \equiv \dfrac{\mean{\Sigma_u}}{\Lambda_f}\,.
\label{dim5Y}
\end{equation}
An alternative realization, that we also explore below, is that of a $d=6$ Yukawa operator, involving generically
two scalar flavons for each spurion,
\begin{eqnarray}
\mathscr{L}_Y=\ov{Q}_L\frac{\chi_d^L\chi_d^{R\dag}}{\Lambda_f^2}D_RH+\ov{Q}_L\frac{\chi_u^L\chi_u^{R\dag}}
{\Lambda_f^2}U_R\tilde{H}+\text{h.c.}\,,
\label{YukLagrangian_Fund}
\end{eqnarray}
which provide the following relations between Yukawa couplings and vevs:
\begin{eqnarray}
Y_D \equiv \dfrac{\mean{\chi_d^L}\mean{\chi_d^{R\dag}}}{\Lambda_f^2}\,, \qquad \qquad
Y_U \equiv \dfrac{\mean{\chi_u^L}\mean{\chi_u^{R\dag}}}{\Lambda_f^2}\,.
\label{YukawasVEVs_Bi_Fund_2F}
\end{eqnarray}
In this interesting case, the flavons are simply vectors under the flavour group, alike to quarks, with the
simplest quantum number assignment being $\chi_{u,d}^L\sim(3,1,1)$, $\chi_u^R\sim(1,3,1)$ and
$\chi_d^R\sim(1,1,3)$. Following this pattern, would the Yukawa couplings result from a condensate of fermionic
flavons \cite{CG:MFV}, a $d=7$ Yukawa operator could be adequate
\begin{equation}
Y_D \equiv \dfrac{\mean{\ov{\Psi}_d^L\Psi_d^R}}{\Lambda_f^3}\,,\qquad \qquad
Y_U \equiv \dfrac{\mean{\ov{\Psi}_u^L\Psi_u^R}}{\Lambda_f^3}\,,
\label{dim7Y}
\end{equation}
with fermions quantum numbers under $G_f$ as in the previous case.
Notice that these realizations in which the Yukawa couplings correspond to the vev of an aggregate of fields,
rather than to a single field, are not the simplest realization of MFV as defined in Ref.~\cite{DGIS:MFV},
while still corresponding to the essential idea that the Yukawa spurions may have a dynamical origin.
The goal of this work is to address the problem of the determination of the general scalar potential, compatible
with the flavour symmetry $G_f$, for the flavon fields denoted above by $\Sigma$ or $\chi$.
An interesting question is whether it is possible to obtain the SM Yukawa pattern - i.e. the observed values of quark
masses and mixings- with a renormalizable potential.
We derive the potential, analyze the possible vacua, and discuss the degree of ``naturalness'' of the possible solutions.
It will be shown that the possibility of
obtaining a large mass hierarchy and mixing at the renormalizable level varies much depending on the dimension of
the Yukawa operator. The role played by non--renormalizable terms and the fine--tunings required to
accommodate the full spectrum will be explored.
A relevant issue is what will be meant by natural:
following 't Hooft's naturalness
criteria, all dimensionless free parameters of the potential not
constrained by the symmetry should be of order one, and all dimensionful ones are expected to be of the order of the scale(s) of the
theory. We will thus explore in which cases --if any-- those criteria allow that the minimum of the MFV potential
corresponds automatically to mixings and large mass hierarchies. Stronger than \cal{O}($10$\%) adjustments (typical
Clebsh-Gordan values in any theory)
will be considered fine-tuned.
It is worth to note that the structure of the scalar potentials constructed here is more general than the particular
effective realization in eqs.~(\ref{LagrangianBi_Fund}) and (\ref{YukLagrangian_Fund}). Indeed, it relies exclusively
on invariance under the symmetry $G_f$ and on the flavon representation, bi-fundamental or fundamental{\footnote{For
instance, the potential and the consequences for mixing obtained in this work will apply as well to the construction
in ref.~\cite{GRV:SU3gauged}, notwithstanding the fact that there the flavon vevs show and inverse hierarchy than that for the minimal version of MFV, as they are proportional to the inverse of the SM Yukawa couplings.}
We limit our detailed discussions below to the quark sector. The implementation of MFV in the leptonic
sector~\cite{CGIW:MLFV,DP:MLFV} requires some supplementary assumptions, as Majorana neutrino masses require
to extend the SM and involve a new scale: that of lepton number violation. Due to the smallness of neutrino
masses, the {\it effective} scale of lepton number violation must be distinct from the flavour and electroweak
ones, if new observable flavour effects are to be expected~\cite{GHHH:MFSM}.
Nevertheless, the analysis of the flavon scalar potential performed below may also apply when considering
leptons, although the precise analysis and implications for the leptonic spectrum will be carried
out elsewhere.
The structure of the manuscript is as follows. In sect.~\ref{sec:2F}, for the two-family case we analyze the
renormalizable potential for $d=5$ and $d=6$ Yukawa operators, or in other words of flavons in the bi-fundamental
and in the fundamental of $G_f$, respectively, showing that in the latter case mixing and a strong hierarchy are
intrinsically present. The corrections induced by non-renormalizable terms are also discussed. In sect.~\ref{sec:3F}
the analogous analyses are carried out for the realistic three-family case and it is also discussed the qualitative new
features appearing when considering simultaneously $d=5$ and $d=6$ Yukawa operators. The conclusions are
presented in sect.~\ref{sec:concl}. Details of the analytical and numerical discussions of the potential minimization
can be found in the Appendices.
\section{Two Family Case}
\label{sec:2F}
We start the discussion of the general scalar potential for the MFV framework by illustrating the two-family case,
postponing the discussion of three families to the next section. Even if we restrict to a simplified case, with a smaller
number of Yukawa couplings and mixing angles, it is a very reasonable starting-up scenario, that corresponds to the
limit in which the third family is decoupled, as suggested by the hierarchy between quark masses and the smallness
of the CKM mixing angles\footnote{We follow in this paper the PDG~\cite{PDG2010} conventions for the CKM matrix
parametrization.} $\theta_{23}$ and $\theta_{13}$. In this section, moreover, we will introduce most of the conventions
and ideas to be used later on for the three-family analysis.
With only two generations the non-Abelian flavour symmetry group, $G_f$, is reduced to
\begin{equation}
G_f=SU(2)_{Q_L}\times SU(2)_{U_R}\times SU(2)_{D_R}\,,
\end{equation}
under which the quark fields transform as
\begin{equation}
Q_L\sim(2,1,1)\,,\qquad\qquad
U_R\sim(1,2,1)\,,\qquad\qquad
D_R\sim(1,1,2)\;.
\end{equation}
Following the MFV prescription, in order to preserve the flavour symmetry in the Lagrangian, the Yukawa spurions
introduced in eq.~(\ref{spurions}) now transform under $G_f$ as
\begin{equation}
Y_U\sim(2,\ov{2},1)\,,\qquad\qquad
Y_D\sim(2,1,\ov{2})\,.
\end{equation}
The masses of the first two generations and the mixing angle among them arise once the spurions take the following values:
\begin{eqnarray}
Y_D=\left(
\begin{array}{cc}
y_d & 0 \\
0 & y_s \\
\end{array}
\right)\;,\qquad & & \qquad
Y_U={\mathcal V}_C^\dag\left(
\begin{array}{cc}
y_u & 0 \\
0 & y_c \\
\end{array}\right)\,, \label{SpurionsVEVs2F}
\end{eqnarray}
where
\begin{eqnarray}
{\mathcal V}_C =\left(
\begin{array}{cc}
\cos\theta & \sin\theta \\
-\sin\theta & \cos\theta \\
\end{array}
\right)
\label{Cabibbo}
\end{eqnarray}
is the usual Cabibbo rotation among the first two families.
\mathversion{bold}
\subsection{$d=5$ Yukawa operators: the bi-fundamental approach}
\label{sec:Bi_Fund2F}
\mathversion{normal}
The most intuitive approach, in looking for a dynamical origin of MFV, is probably to promote each Yukawa coupling
from a simple spurion to a flavon field. In other words, to consider the effective $d=5$ Lagrangian described
above in eq.~(\ref{LagrangianBi_Fund}). The new fields --flavons-- are singlets under the SM gauge group but have, for the
two-family case, the non-trivial transformation properties under $G_f$ given by
\begin{equation}
\Sigma_u\sim(2,\ov{2},1)\longrightarrow\Sigma_u'=\Omega_L\, \Sigma_u\, \Omega_{U_R}^\dag\;,\qquad
\Sigma_d\sim(2,1,\ov{2})\longrightarrow\Sigma_d'=\Omega_L\, \Sigma_d\, \Omega_{D_R}^\dag\,,
\label{FlavonsTransformations2F_Bifund}
\end{equation}
where $\Omega_X$ denotes the doublet transformation under the $SU(2)_X$-component of the flavour group. Once these
flavon fields develop vevs as in eq.~(\ref{dim5Y}) and eq.~(\ref{SpurionsVEVs2F}), the flavour symmetry is explicitly
broken and quark masses and mixings are originated. The effective field theory obtained at the electroweak scale is
exactly MFV \cite{DGIS:MFV} (restricted to the two-family case). Then, within this approach, the problem of the origin
of flavour is replaced by the need to explain if and how this particular vev configuration can naturally arise from
the minimization of the associated scalar potential.
This minimal framework can be easily extended in different ways, such as, for instance:
\begin{itemize}
\item Considering different scales for the $\Sigma_u$ and $\Sigma_d$ flavon vevs.
\item Adding new representations. The most straightforward way to complete the basis in
eq.~(\ref{FlavonsTransformations2F_Bifund}), is to add a third flavon transforming as a bi-fundamental
of the RH components:
\begin{equation}
\Sigma_R \sim (1,2,\ov2) \longrightarrow \Sigma_R'=\Omega_{U_R}\, \Sigma_R\, \Omega_{D_R}^\dag\,.
\label{SigmaR}
\end{equation}
This new field does not contribute to the Yukawa terms, at least at the renormalizable level, but introduces new
operators with respect to MFV, which induce flavour changing neutral currents (FCNC) mediating fully right-handed (RH)
processes\footnote{The phenomenological impact of these operators has already been introduced and studied in the
three-family case in ref.~\cite{BGI:MFVwithRHcurrents}, in a different context.}.
\item Adding new replicas of the bi-fundamental representations.
This could be very helpful as a natural source of new scales and possible mixings.
\end{itemize}
The first two possibilities do not affect essentially the flavour structure of the quark Yukawa couplings, which
is the focus of this work, and we will not consider them below. No further consideration is given either in this section
to the third possibility, both for the sake of simplicity and because of the aesthetically unappealing aspect of being
a trivial replacement of the puzzle of quark replication with that of flavon replication.
We will thus restrict the remaining of this section to the analysis of the potential for just one $\Sigma_u$ and one
$\Sigma_d$ fields, eqs.~(\ref{FlavonsTransformations2F_Bifund}). The general scalar potential, can then be written as a
sum of two parts, the first dealing only with the SM Higgs fields and the second accounting also for the flavons interactions:
\begin{equation}
V\equiv V_H+V_\Sigma=-\mu^2 H^\dag H+\lambda_H (H^\dag H)^2+\sum_{i=4}^\infty V^{(i)}[H,\Sigma_u,\Sigma_d]\,.
\label{ScalarPotential}
\end{equation}
Inside $V^{(i)}$ all possible scalar potential terms of the effective field theory are included. In particular,
$V^{(4)}$ contains all the renormalizable couplings written in terms of $H$ and $\Sigma_{u,d}$ while $V^{(i>4)}$
incorporate all the non-renormalizable higher dimensional operators. There is no particular reason to impose that the
EW and the flavour symmetry breaking should occur at the same scale. Indeed it is plausible that the flavour symmetry
is broken by some new physics mechanism at a larger energy scale. Although it is true that the mixed Higgs-flavons terms
could affect the value and location of the electroweak and flavour minima, the flavour composition of each term will not
be modified by them. Once the flavour symmetry breaking occurs, all the terms in $V^{(i)}$ either contribute to the
scalar potential as constants or can be redefined into $\mu^2$ or $\lambda_H$. In what follows the analysis is
restricted to consider only the flavon part of the scalar potential, $V^{(i)}[\Sigma_u,\Sigma_d]$.
\subsubsection{The Scalar Potential at the Renormalizable Level}
From the transformation properties in eq.~(\ref{FlavonsTransformations2F_Bifund}), it is straightforward to write
the most general independent invariants that enter in the scalar potential. At the renormalizable level, and for
the case of two generations, five independent invariants can be constructed\footnote{Any other invariant operator
can be expressed in terms of these five independent invariants. For example: $Tr\left(\Sigma_u\Sigma_u^\dagger
\Sigma_u\Sigma_u^\dagger\right)=Tr\left(\Sigma_u\Sigma_u^\dagger\right)^2-2\det\left(\Sigma_u\right)^2$.}
\cite{Feldmann:2009dc}:
\begin{equation}
\begin{array}{ll}
A_u=\text{Tr}\left(\Sigma_u\Sigma_u^\dagger\right)\,,&\qquad B_u=\det\left(\Sigma_u\right)\,,\\[2mm]
A_d=\text{Tr}\left(\Sigma_d\Sigma_d^\dagger\right)\,,& \qquad B_d=\det\left(\Sigma_d\right)\,,\\[2mm]
A_{ud}=\text{Tr}\left(\Sigma_u\Sigma_u^\dagger\Sigma_d\Sigma_d^\dagger\right)\,.&
\end{array}
\label{Invariants2F_BiFund}
\end{equation}
Eqs. (\ref{dim5Y}) and (\ref{SpurionsVEVs2F}) allow to express these invariants in terms of physical
observables, i.e. the four Yukawa eigenvalues and the Cabibbo angle:
\begin{eqnarray}
\mean{\Sigma_d} = \Lambda_f\,\left(
\begin{array}{cc}
y_d & 0 \\
0 & y_s \\
\end{array}
\right)\;,\qquad & & \qquad
\mean{\Sigma_u}=\Lambda_f\,{\mathcal V}_C^\dag\left(
\begin{array}{cc}
y_u & 0 \\
0 & y_c \\
\end{array}\right)\,,
\label{SigmaSpurionsVEVs2F}
\end{eqnarray}
leading to:
\begin{equation}
\begin{array}{l}
\mean{A_u}=\Lambda_f^2\, (y_u^2+y_c^2)\,,\qquad \qquad \qquad \mean{B_u}=\Lambda_f^2\, y_u\,y_c\,,\\[2mm]
\mean{A_d}=\Lambda_f^2\, (y_d^2+y_s^2)\,,\qquad \qquad \qquad \mean{B_d}=\Lambda_f^2\, y_d\,y_s\,,\\[2mm]
\mean{A_{ud}}=\Lambda_f^4 \left[\left(y_c^2-y_u^2\right)\left(y_s^2-y_d^2\right)\cos2\theta+
\left(y_c^2+y_u^2\right)\left(y_s^2+y_d^2\right)\right]/2\,.
\end{array}
\label{InvariantsExplicit2F_BiFund}
\end{equation}
Notice that the mixing angle appears only in the vev of $A_{ud}$, which is the only operator that mixes
the up and down flavon sectors. This is as intuitively expected: {\it the mixing angle describes the relative
misalignment between two different directions in flavour space}. It is also interesting to notice that the expression
for $\mean{A_{ud}}$ is related to the Jarlskog invariant for two families,
\begin{equation}
4J=4\det{\left[Y_UY_U^\dagger,Y_DY_D^\dagger\right]}=\left(\sin{2\theta}\right)^2
\left(y_c^2-y_u^2\right)^2\left(y_s^2-y_d^2\right)^2 \,,\nonumber
\end{equation}
by the following relation:
\begin{equation}
\dfrac{1}{\Lambda_f^4}\dfrac{\partial \mean{A_{ud}}}{\partial \theta}= -2 \sqrt{J}\,.
\end{equation}
Using the invariants in eqs.~(\ref{Invariants2F_BiFund}), the most general renormalizable scalar
potential allowed by the flavour symmetry reads:
\begin{equation}
V^{(4)}=\sum_{i=u,d}\left(-\mu_i^2A_i-\tilde{\mu}_i^2B_i+\lambda_iA_i^2+
\tilde{\lambda}_i B_i^2\right)+g_{ud}A_uA_d+f_{ud}B_uB_d+\sum_{i,j=u,d}h_{ij}A_iB_j+\lambda_{ud}A_{ud}\,,
\label{ScalarPot2F_BiFund}
\end{equation}
where strict naturalness criteria would require all dimensionless couplings $\lambda$, $f$, $g$, $h$ to be of
order $1$, and the dimensionful $\mu$-terms to be
smaller or equal than $\Lambda_f$ although of the same order of magnitude. It is clear from the start that, with the only
use of symmetry implemented here, a strict implementation of such criteria could lead at best to a strong hierarchy with
some fields massless and the rest with masses of about the same scale. The ``fan" structure of quark mass splittings
observed clearly calls, instead, for a readjustment of the relative size of some $\mu$ parameters, at least when
restraining to the analysis of the renormalizable and classic terms of the potential. One question is whether, in this
situation, even further fine-tunings are required among the mass parameters in the potential to accommodate nature.
The relations in eq.~(\ref{InvariantsExplicit2F_BiFund}) allow
to determine the positions of the potential minima in terms of physical observables. A careful analytical and
numerical study of the potential can be found in the Appendices. Here we briefly comment on the most relevant
physical results. Consider first the angular part of the potential. Deriving $V^{(4)}$ with
respect to the angle $\theta$, it follows that
\begin{equation}
\left.\dfrac{\partial V^{(4)}}{\partial \theta}\right|_{min}\equiv \lambda_{ud}\dfrac{\partial{\mean{A_{ud}}}}{\partial \theta}\propto
\lambda_{ud}\sin{2\theta}\left(y_c^2-y_u^2\right)\left(y_s^2-y_d^2\right) \propto \lambda_{ud} \sqrt{J}\,.
\label{CabibboEq2F_BiFund}
\end{equation}
The minimum of the scalar potential thus occurs when at least one of the following conditions is satisfied
i) $\lambda_{ud}=0$, ii) $\sin\theta=0$, iii) $\cos\theta=0$ or iv) two Yukawas in the same sector are degenerate.
When condition i) is imposed, the angle remains undetermined;
this assumption corresponds however to a severe fine-tuning on the model, as no symmetry protects this term
from reappearing at the quantum level. Instead, due to the smallness of the Cabibbo angle, condition ii) can be
interpreted as a first order solution which needs to be subsequently corrected, for example by the introduction
of higher order operators. This possibility will be discussed in more detail in the next subsection. Finally, the last
conditions, iii) and iv), are phenomenologically non representative of nature and large (higher order) corrections
should be advocated in order to diminish the angle or to split the Yukawa degeneracy, respectively, making these
solutions unattractive. All in all, the straightforward lesson that follows from eq.~(\ref{CabibboEq2F_BiFund})
is that, given the mass splittings observed in nature, {\it the scalar potential for bi-fundamental flavons does
not allow mixing at leading order.}
From the requirement that the derivatives of the scalar potential with respect to $y_{u,d,c,s}$ also vanish at the
minima, four additional independent relations on the physical parameters are obtained. As discussed above, to obtain simultaneously a sizeable mixing and a mass spectrum largely splitted in masses, instead of generically degenerate, it is necessary to (re-)introduce a large, and unnatural, hierarchy among the
different operators appearing in the scalar potential (see Appendix B for numerical details).
These observations can be summarized stating that, with a natural choice of the coefficients appearing in the
renormalizable scalar potential $V^{(4)}$, after minimization one naturally ends up with a vanishing or undetermined
mixing angle and with a naturally degenerate spectrum.
In this respect we agree with a remark that can be found in refs.~\cite{Feldmann:2009dc,GRV:SU3gauged}. It is, however,
interesting to notice that if the invariants $B_{u,d}$ (i.e. the determinant of the flavons) are neglected, which could
be justified for example introducing some {\it ad hoc} discrete symmetry, the minima equations would then allow, instead,
solutions non-degenerate in mass for same-charge quarks, with non-vanishing Yukawa couplings for the first (second)
quark generations. This may open the possibility to study a modified version of the scalar potential in eq.~(\ref{ScalarPot2F_BiFund}), that predicts a natural
hierarchy among the Yukawas of different generations.
\subsubsection{The Scalar Potential at the Non-Renormalizable Level}
Consider the addition of non-renormalizable operators to the scalar potential, $V^{(i>4)}$. It is very interesting to
notice that this does {\it not} require the introduction of new invariants beyond those in eq.~(\ref{Invariants2F_BiFund}):
all higher order traces and determinants can in fact be expressed in terms of that basis of five ``renormalizable''
invariants.
The lowest higher dimensional contributions to the scalar potential have dimension six (the complete list can be
found in Appendix A). At this order, the only terms affecting the mixing angle are
\begin{equation}
V^{(6)}\supset \dfrac{1}{\Lambda_f^2}\sum_{i=u,d}\left(\alpha_iA_{ud}B_i+\beta_iA_{ud}A_i\right)\,.
\end{equation}
These terms, however, show the same dependence on the Cabibbo angle previously found in eq.~(\ref{CabibboEq2F_BiFund})
and, consequently, they can simply be absorbed in the redefinition of the lowest order parameter, $\lambda_{ud}$.
In other words, even at the non-renormalizable level, the most favorable trend leads to no mixing. To find a
non-trivial angular structure it turns out that terms in the potential of dimension eight (or higher) have to be
considered, that is
\begin{equation}
V^{(8)}\supset \lambda_{udud}A_{ud}^2\,,
\end{equation}
and eq.~(\ref{CabibboEq2F_BiFund}) would be replaced by
\begin{equation}
\left.\dfrac{\partial V}{\partial \theta}\right|_{min}\propto \sin{2\theta}\left(y_c^2-y_u^2\right)\left(y_s^2-y_d^2\right)
\left[\lambda_{ud}-2\,y^2_c y^2_s \lambda_{udud} \sin^2\theta + \dots \right]\,,
\end{equation}
implying
\begin{equation}
\sin^2\theta\simeq \dfrac{\lambda_{ud}}{2\, y^2_c y^2_s \lambda_{udud}}\,.
\end{equation}
Using the experimental values of the Yukawa couplings $y_s$ and $y_c$, a meaningful value for $\sin\theta$ can be
obtained although at the price of assuming a highly fine-tuned hierarchy between the dimensionless coefficients of $d=4$ and
$d=8$ terms, $\lambda_{ud}/\lambda_{udud}\sim 10^{-10}$, that cannot be naturally justified in an effective Lagrangian approach.
The remaining four equations defining the minima, obtained deriving the scalar potential with respect to $y_{u,d,c,s}$,
lead to no improvement as compared to the renormalizable case: the Yukawa couplings are always given by general
combinations of the coefficients of the scalar potential, underlining the complete absence of hierarchies among them.
Realistic masses can be obtained at the classical level only when suitable fine-tunings are enforced~\footnote{See note added in proof}.
To summarize, it is possible to account for a non-vanishing mixing angle adding non-renormalizable terms to the scalar
potential, although at the prize of introducing a large fine-tuning. This requirement comes in addition to the fact that
the hierarchies among the Yukawa couplings can only be imposed by hand. Therefore the use of bi-fundamental scalar
fields leads to an unsatisfactory answer to the problem of explaining the origin of flavour within the MFV hypothesis.
For the sake of illustrating the argument with a practical exercise, we conclude this section showing, as an explicit
example, a fine-tuned scalar potential which can allow hierarchical Yukawas and a non-vanishing mixing angle:
\begin{equation}
V=\sum_i\left(-\mu_i^2A_i+\tilde{\lambda}_iB_i^2+\lambda_iA_i^2\right)+
\dfrac{\lambda_{udud}}{\Lambda_f^4}\left(A_{udud}-2A_{uudd}\right)- \epsilon_b \tilde{\mu}_d^2 B_d-
\epsilon_u \tilde{\mu}_u^2 B_u+ \epsilon_\theta\lambda_{ud}A_{ud}\,,
\label{finetunepot}
\end{equation}
where $\epsilon_{u,d,\theta}$ are suppressing factors,
which could justified via some discrete symmetry, and $A_{uudd}$, $A_{udud}$, dimension eight invariants defined by the following relations:
\begin{equation}
A_{udud}=\text{Tr}\left(\Sigma_u \Sigma^\dag_u \Sigma_d \Sigma^\dag_d \Sigma_u \Sigma^\dag_u \Sigma_d \Sigma_d^\dag\right)\,,\qquad
A_{uudd}=\text{Tr}\left(\Sigma_u \Sigma^\dag_u \Sigma_u \Sigma^\dag_u \Sigma_d \Sigma^\dag_d \Sigma_d \Sigma_d^\dag\right)\,.
\label{Invariants2F_BiFund1}
\end{equation}
By minimizing the potential in eq.~(\ref{finetunepot}) one obtains the following values for the Yukawa eigenvalues and
the Cabibbo angle:
\begin{equation}
\begin{gathered}
y_u \simeq \epsilon_u\,\dfrac{\sqrt{\lambda_u}\,\tilde\mu_u}{\sqrt2\,\tilde\lambda_u\,\mu_u}
\dfrac{\tilde\mu_u}{\Lambda_f}\,,\qquad\qquad
y_d \simeq \epsilon_d\, \dfrac{\sqrt{\lambda_d}\,\tilde\mu_d}{\sqrt2\,\tilde\lambda_d\,\mu_d}
\dfrac{\tilde\mu_d}{\Lambda_f}\,,\\[4mm]
y_c \simeq \dfrac{\mu_u}{\sqrt2\,\Lambda_f\,\sqrt\lambda_u}\,,\qquad\qquad
y_s \simeq \dfrac{\mu_d}{\sqrt2\,\Lambda_f\,\sqrt\lambda_d}\,,\\[4mm]
\sin^2\theta \simeq \epsilon_\theta\,\dfrac{\lambda_{ud}}{\lambda_{udud}\,y_c^2\,y_s^2}\,.
\end{gathered}
\end{equation}
Imposing for no good reason the values $\epsilon_u \sim 10^{-3}$, $\epsilon_d\sim5\times10^{-2}$, $\epsilon_\theta
\sim10^{-10}$ and $\mu/(\sqrt{\lambda}\Lambda_f)\approx\tilde\mu/(\sqrt{\tilde\lambda}\Lambda_f)\sim 10^{-3}$, the
correct hierarchies between the quark masses and the correct Cabibbo angle could be obtained (see details for this
special case in Appendix B).
The discussion about $d=8$ terms presented above has pure illustrative purposes, as it may be a priori misleading to discuss
the effects of $d=8$ terms in the potential without simultaneously considering quantum or other higher-order sources
of corrections, such as the possible impact of a $\Sigma_R$ flavon\footnote{The impact of the fully RH bi-fundamental
$\Sigma_R$ is negligible: indeed it can enter in the scalar potential only as powers of $\Sigma_R\Sigma_R^\dag$ or
its hermitian conjugate, and in particular, being a singlet of $SU(2)_{Q_L}$, it cannot mix with the other flavons.
As a result, its contributions can always be absorbed through a redefinition of the parameters and then the conclusions
above still hold.} - see eq.~(\ref{SigmaR}) -} or other $G_f$ representations.
\mathversion{bold}
\subsection{$d=6$ Yukawa operator: the fundamental approach}
\label{sec:Fund2F}
\mathversion{normal}
The identification of the Yukawa spurions as single flavon fields, transforming in the bi-fundamental representation
of the flavour group (e.g. for a $d=5$
Yukawa operator), is only one of the possible ways the MFV ansatz can be implemented. An attractive alternative is to consider the Yukawas as composite objects or aggregates of several
fields, e.g. suggesting Yukawa operators with $d>5$. In the simplest case, each Yukawa corresponds to two scalar fields $\chi$ transforming in the
fundamental representation of $G_f$ (e.g. $Y\sim \mean{\chi} \mean{\chi'^\dagger}/\Lambda_f^2$, see
eqs.~(\ref{YukLagrangian_Fund}) and (\ref{YukawasVEVs_Bi_Fund_2F})). This approach would {\it a priori} allow to introduce one
new field for each component of the flavour symmetry: i.e. to reconstruct the spurions in eq.~(\ref{spurions}) just out
of three vectors transforming as $(2,1,1)$, $(1,2,1)$ and $(1,1,2)$. However, such a minimal setup leads to an
unsatisfactory realization of the flavour sector as no physical mixing angle is allowed at the renormalizable level~\footnote{
Because then the flavons associated to the up and down left-handed character are not misaligned in flavour space, but correspond instead to
just one $(2,1,1)$ flavon.}. The situation improves qualitatively, though, if two $(2,1,1)$ representations are introduced,
one for the up and one for the down quark sectors. Consider then the following four fields:
\begin{equation}
\chi_u^L \in \left( 2, 1, 1\right)\,,\qquad\quad
\chi_u^R \in \left( 1, 2, 1\right)\,,\qquad\quad
\chi_d^L \in \left( 2, 1, 1\right)\,,\qquad\quad
\chi_d^R \in \left( 1, 1, 2\right)\,.
\label{listfundamentals}
\end{equation}
The corresponding $d=6$ effective Lagrangian and Yukawa couplings have been shown in eqs.~(\ref{YukLagrangian_Fund})
and (\ref{YukawasVEVs_Bi_Fund_2F}). These flavons are vectors under the flavour symmetry. The only physical
invariants that can be associated to vectors are the norm of the vectors and, eventually, their relative angles.
Any matrix resulting from multiplying two vectors has only one non-vanishing eigenvalue, independently of the number
of dimensions of the space. This fact alone already implies that, at the leading renormalizable order under discussion,
just one ``up"-type quark and one ``down"-type quark are massive: a strong mass hierarchy between quarks of the same
electric charge is thus automatic in this setup, which is a very promising first step in the path to explain the
observed quark mass hierarchies.
More in detail, the resulting Yukawa matrices are general $2\times2$ matrices, containing many unphysical parameters. Without loss
of generality, it is possible to express the Yukawa couplings in terms of physical quantities by choosing the flavon
vevs as follows:
\begin{equation}
{\mean{\chi_i}}\equiv\left|\chi_i \right| {\mathcal V}_i
\left(
\begin{array}{c}
0 \\ 1
\end{array}
\right)\,,
\label{VEV2F_Fund}
\end{equation}
where by $|\chi_i |$ we denote the norm of the vev of $\chi$, $|\chi_i | \equiv|\mean{\chi_i }|$, and ${\mathcal V}_i$
are $2\times2$ unitary matrices.
Redefining the quark fields as follows,
\begin{equation}
Q_L'={\mathcal V}_L^{(d)\dag} Q_L\,,\qquad
U_R'={\mathcal V}_R^{(u)\dag} U_R\,,\qquad
D_R'={\mathcal V}_R^{(d)\dag} D_R\,,
\end{equation}
it results
\begin{equation}
\mathscr{L}_Y=\overline{Q}_L' Y_D D_R' H+ \overline{Q}_L' Y_U U_R' \tilde{H} +\text{h.c.}\,,
\end{equation}
with the corresponding Yukawa matrices given by~\footnote{The
cutoff scale $\Lambda_f$ refers to the scale of the flavour dynamics. In principle we could have different scales
for the left and right flavons as well as for the up and down ones, but for simplicity we assume that all the
scales are close and $\Lambda_f$ refers to the average value.}
\begin{equation}
Y_D=\dfrac {\left|\chi_d^L\right|\left|\chi_d^R\right|}{\Lambda_f^2} \left(\begin{array}{cc}
0 & 0\\
0 & 1
\end{array}\right)\,,\qquad\quad
Y_U=\dfrac{\left|\chi_u^L\right|\left|\chi_u^R\right|}{\Lambda_f^2}\, {\mathcal V}_L^{(d)\dag} {\mathcal V}_L^{(u)}
\left(\begin{array}{cc}
0&0\\
0&1\end{array}\right)\,.
\label{Yukawas2F_Fund}
\end{equation}
This illustrates explicitly that: i) {\it there is a natural hierarchy among the mass of the first
and second generations, without imposing any constraint on the parameters of the scalar potential}; ii) {\it the product
${\mathcal V}_L^{(d)\dag} {\mathcal V}_L^{(u)}$ is a non-trivial unitary matrix that contains all the information about
the mixing angle} (the phase can be easily removed in the two-family case under discussion).
There is now a clear geometrical
interpretation of the Cabibbo angle: {\it the mixing angle between two generations of quarks is the misalignment
of the $\chi^L$ flavons in the flavour space}, with the mixing matrix appearing in weak currents, eq.~(\ref{Cabibbo}), given by
\begin{equation}
{\mathcal V_C}={\mathcal V}_L^{(u)\dag}\, {\mathcal V}_L^{(d)}\,.
\end{equation}
Let us compare the phenomenology expected from bi-fundamental flavons (i.e. $d=5$ Yukawa operator) with that from fundamental
flavons (i.e. $d=6$ Yukawa operators). For bi-fundamentals, the list of effective FCNC operators is exactly the same that
in the original MFV proposal~\cite{DGIS:MFV}. The case of fundamentals presents some differences: higher-dimension invariants
can be constructed in this case, exhibiting lower dimension than in the bi-fundamental case.
For instance, one can compare these two operators:
\begin{equation}
\ov{D}_R\,\Sigma_d^\dag\,\Sigma_u\,\Sigma_u^\dag\,Q_L\sim[\text{mass}]^6\qquad \longleftrightarrow \qquad\ov{D}_R\,
\chi_d^{R}\,\chi_u^{L\dag}\,Q_L\sim[\text{mass}]^5\,,
\label{OperatorsFCNCdimensions}
\end{equation}
where the mass dimension of the invariant is shown in brackets; with these two types of basic bilinear FCNC
structures it is possible to build effective operators describing FCNC processes, but differing on the degree of
suppression that they exhibit.
This underlines the fact that the identification of Yukawa couplings with aggregates of two or more flavons is a
setup which goes technically beyond the realization of MFV, resulting possibly in a distinct phenomenology which
could provide a way to distinguish between fundamental and bi-fundamental origin.
\subsubsection{The scalar potential}
\label{sec:Fund2F_ScalarPot}
The general scalar potential that can be written including flavons in the fundamental is analogous to that in eq.~(\ref{ScalarPotential}), replacing $\Sigma_i$ with $\chi_i$,
\begin{equation}
V\equiv V_H+V_\chi\,.
\end{equation}
Previous considerations regarding the scale separation between EW and flavour breaking scale hold also in this case,
and in consequence the Higgs sector contributions will not be explicitly described.
Any flavour invariant operator can be constructed out of the following five independent
building blocks:
\begin{equation}
\chi_u^{L\dag}\chi_u^L\,,\qquad\qquad
\chi_u^{R\dag}\chi_u^R\,,\qquad\qquad
\chi_d^{L\dag}\chi_d^L\,,\qquad\qquad
\chi_d^{R\dag}\chi_d^R\,,\qquad\qquad
\chi_u^{L\dag}\chi_d^L\,.
\end{equation}
From the expressions for the Yukawa matrices in eqs.~(\ref{Yukawas2F_Fund}), it follows that in this scenario the scalar
potential depends only on three of the five physical parameters: one angle and the two (larger) Yukawa couplings
\begin{equation}
\left|\chi_u^L\right|\left|\chi_u^R\right|=\Lambda_f^2 \, y_c\,,\qquad \qquad \,
\left|\chi_d^L\right|\left|\chi_d^R\right|=\Lambda_f^2 \, y_s\,,\qquad \qquad \,
\chi_u^{L\dagger}\chi_d^L=\cos\theta_c\left|\chi_u^L\right|\left|\chi_d^L\right|\,,
\label{connectionfund}
\end{equation}
given by the product of the left and right up (down) flavon moduli. As expected, the mixing angle is simply
the angle defined in flavour space by the up and down left vectors. From the point of view of the measurable quantities,
there is a certain parametrization freedom, and a possible convenient choice is given by\footnote{See Appendix C for
a detailed discussion.}
\begin{equation}
\frac{\left|\chi_u^R\right|}{\Lambda_f}=\,1\,=\frac{\left|\chi_d^R\right|}{\Lambda_f}\,.
\label{RHvevs}
\end{equation}
As a result, the invariants physically relevant for the flavour structure are:
\begin{equation}
\left|\chi_u^L\right|=\Lambda_f \,y_c\,,\qquad\quad
\left|\chi_d^L\right|=\Lambda_f \, y_s\,,\qquad\quad
\chi_u^{L\dag}\chi_d^L=\Lambda_f^2 y_c\,y_s\,\cos{\theta}\,.
\label{vevsFundamentals2D}
\end{equation}
At the renormalizable level, the scalar potential is given by
\begin{equation}
V^{(4)}=-\sum_{i=u,d}\mu_i^2\, \chi_i^{L\dag}\chi_i^L-
\sum_{i=u,d}\tilde{\mu}_i^2\,\chi_i^{R\dag}\chi_i^R -
\mu_{ud}^2\,\chi_u^{L\dag}\chi_d^L+\ldots\,,
\end{equation}
where dots stand for all possible quartic couplings. The total number of operators that can be introduced at
the renormalizable level is 20. However, as shown in Appendix C, many of them (i.e. quartic couplings that mix
different flavours) do not have any real impact on the existence and determination of the minima.
Studying the latter, the following relations
between the (large) up and down Yukawa eigenvalues and the Cabibbo angle follow:
\begin{equation}
\frac{y_{s}^2}{y_{c}^2}=\frac{\mu_d^2\lambda_u}{\mu_u^2\lambda_d}\, , \qquad \qquad \cos\theta=
\frac{\sqrt{\lambda_u\lambda_d}\mu_{ud}^2}{\lambda_{ud}\mu_u\mu_d}\,
\label{YukawasResults}
\end{equation}
which shows that without strong fine-tunings this scenario can explain the hierarchy between the first and second
family, and account for a sizable Cabibbo angle.
\subsubsection{The first generation}
\label{sec:Fund2F_FirstGen}
In this two-generation analysis, the first family has remained massless at the renormalizable level. A first
possibility is that non-renormalizable corrections may induce this small masses.
Non-renormalizable interactions manifest themselves in form of higher order contributions to the Yukawa operators
and the flavon vevs and/or as non-renormalizable terms in the potential, which can modify its minima.
From eq.~(\ref{YukLagrangian_Fund}) and the flavon transformation properties, it follows that higher order
contributions to the Yukawa operators can only be constructed by further insertions of $\chi^\dag\chi$ inside the
renormalizable operators. However, such kind of insertions do not modify the flavour structure of the Yukawa
matrices, but simply redefine the two heavier couplings, $y_c$ and $y_s$.
On the other hand, the introduction of higher order operators in the scalar potential has the effect of modifying
the vevs of the flavons, replacing the relation in eq.~(\ref{VEV2F_Fund}) with
\begin{equation}
\dfrac{\mean{\chi_{u,d}^{L,R}}}{\Lambda_f}\equiv\left|(1+{\mathcal O}(\epsilon))\chi_{u,d}^{L,R}\right|
\left(\mathcal V_{L,R}^{(u,d)}(1+{\mathcal O}(\epsilon) )
\right)
\left(\begin{array}{c}
{\mathcal O}(\epsilon) \\ 1
\end{array}\right)\,,
\end{equation}
where $\epsilon\ll1$ parametrizes the ratio among higher and leading order contributions. The only effect of these
modifications is to redefine the mixing angle $\theta$ and the second family Yukawas, $y_c$ and $y_s$, without
changing the rank of the Yukawa matrices and leaving thus the first generation massless. In summary, non-renormalizable
interactions cannot switch on additional (first family) Yukawas if they were absent at the renormalizable level.
An alternative can be built on the fact that each up-down set of fundamental flavons provides a supplementary scale,
in addition to new sources of mixing from their misalignment. A possibility along this direction is to enlarge
the number of flavons to six, made out of a set of three ($\chi_{u,d}^R$ plus just one $\chi^L$)
replicated: in total
two $\chi^L\sim(2,1,1)$, two $\chi_u^R\sim(1,2,1)$ and two $\chi_d^R\sim(1,1,2)$.
In this case the Yukawa terms change in a non-trivial way:
\begin{equation}
Y_D\equiv\dfrac{\sum_{ij}\alpha^d_{ij}\mean{\chi_i^L}\mean{\chi_j^{R\dag}}}{\Lambda_f^2}\,, \qquad \qquad
Y_U\equiv\dfrac{\sum_{ij}\alpha^u_{ij}\mean{\chi_i^L}\mean{\chi_j^{R\dag}}}{\Lambda_f^2}\,,
\end{equation}
with $\alpha_{i,j}$ numerical coefficients and $i,j$ running over all flavons. An explicit computation reveals that,
for generic values of $\alpha_{ij} (\neq 0)$, the rank of the Yukawa matrices is indeed two. However, in this case,
the natural hierarchy between the first and second family is lost, being all the Yukawas of the same order unless
the vevs of the new flavons are unnaturally smaller than those of the first replica. In conclusion, adding new RH
flavon copies does not lead either to an appealing and natural source of masses for the first generation.
\section{The Three-Family Case}
\label{sec:3F}
Let us extend the previous analysis to the three-family case.
While most of the procedure, with both bi-fundamental and fundamental representations, follows straightforwardly, two main
differences should be underlined. First of all, the top Yukawa coupling, $y_t$, is now a parameter which is of ${\mathcal O}(1)$.
The fact that in the two-family case the largest Yukawa, $y_c$ was much smaller than one, allowed us to safely
retain only the lowest order terms in the (Yukawa) perturbative expansion. In the three-family scenario, in
principle, one should include all orders in the expansion. However, in this case, the Cayley-Hamilton identity~\cite{CNS:SUSYMFVRunning,MS:SUSYMFVEDM} provides a way out, as it proves that
a general $3\times3$ matrix $X$ must satisfy the relation:
\begin{equation}
X^3-\text{Tr}[X]\, X^2+\dfrac{1}{2}X\left(\text{Tr}[X]^2-\text{Tr}\left[X^2\right]\right)-\det[X]=0\,,
\end{equation}
which allows to express all powers $X^n$ (with $n>2$) in terms solely of $\mathbb{1}$, $X$ and $X^2$,
with coefficients involving the traces of $X$ and $X^2$ and the determinant of $X$. In the case under study,
$X$ corresponds to the invariant products $\Sigma^\dag \Sigma$ or $\chi^\dag \chi$, depending on whether bi-fundamental
or fundamental representations are considered.
The second main difference with respect to the two-family case, is the appearance of a physical phase in the
quark mixing matrix. For the sake of simplicity, in this paper we disregard CP-violation,
deferring its discussion to a future work\cite{AGMR:CPpaper}.
\mathversion{bold}
\subsection{$d=5$ Yukawa operator: the bi-fundamental approach}
\label{sec:Bi_Fund3F}
\mathversion{normal}
In this section we extend the approach discussed in sect.~\ref{sec:Bi_Fund2F} to the three-family case.
Consider two bi-triplets under the flavour symmetry $G_f$, see eq.~(\ref{spurions}),
\begin{equation}
\Sigma_u\sim(3,\ov{3},1)\longrightarrow\Sigma_u'=\Omega_L\, \Sigma_u\, \Omega_{U_R}^\dag\;,\qquad
\Sigma_d\sim(3,1,\ov{3})\longrightarrow\Sigma_d'=\Omega_L\, \Sigma_d\, \Omega_{D_R}^\dag\,,
\label{FlavonsTransformations3F_Bifund}
\end{equation}
where now the $\Omega_X$ matrices refer to the triplet transformations under the $SU(3)_X$ component of the flavour
group. The Yukawa Lagrangian is the same as that in eq.~(\ref{LagrangianBi_Fund}). Once the flavons develop a vev
as in eq.~(\ref{dim5Y}), the flavour symmetry is broken and one recovers the observed fermion masses and CKM matrix
given in eq.~(\ref{SpurionsVEVs}).
Recall that the present realization is the simplest realization of the original MFV approach \cite{DGIS:MFV}.
Again, it would be possible to extend it introducing a third RH flavon field, $\Sigma_R\sim (1,3,\ov3)\longrightarrow\Sigma_R'=\Omega_{U_R}\, \Sigma_R\,
\Omega_{D_R}^\dag$. We do not further consider it when constructing the scalar potential, as it cannot contribute
to the Yukawa spurions neither at ${\mathcal O}(1/\Lambda_f)$ nor ${\mathcal O}(1/\Lambda_f^2)$, that is, neither via $d=5$ nor $d=6$
Yukawa operators.
Restricting the explicit analysis to the part of the renormalizable scalar potential not containing the
SM Higgs field, a complete and independent basis is given by the following seven invariant operators:
\begin{equation}
\begin{array}{ll}
A_u=\text{Tr}\left(\Sigma_u\Sigma_u^\dagger\right)\,,& \qquad\qquad
\mean{A_u}=\Lambda_f^2\left(y_t^2+y_c^2+y_u^2\right)\,,\\[2mm]
B_u=\det\left(\Sigma_u\right)\,,& \qquad \qquad
\mean{B_u}=\Lambda_f^3\,y_u\,y_c\,y_t\,,\\[2mm]
A_d=\text{Tr}\left(\Sigma_d\Sigma_d^\dagger\right)\,,& \qquad \qquad
\mean{A_d}=\Lambda_f^2\left(y_b^2+y_s^2+y_d^2\right)\,,\\[2mm]
B_d=\det\left(\Sigma_d\right)\,,& \qquad \qquad
\mean{B_d}=\Lambda_f^3\,y_d\,y_s\,y_b\,,\\[2mm]
A_{uu}=\text{Tr}\left(\Sigma_u\Sigma_u^\dagger\Sigma_u\Sigma_u^\dagger\right)\,,& \qquad \qquad
\mean{A_{uu}}=\Lambda_f^4\left(y_t^4+y_c^4+y_u^4\right)\,,\\[2mm]
A_{dd}=\text{Tr}\left(\Sigma_d\Sigma_d^\dagger\Sigma_d\Sigma_d^\dagger\right)\,,& \qquad \qquad
\mean{A_{dd}}=\Lambda_f^4\left(y_b^4+y_s^4+y_d^4\right)\,,\\[2mm]
A_{ud}=\text{Tr}\left(\Sigma_u\Sigma_u^\dagger\Sigma_d\Sigma_d^\dagger\right)\,,& \qquad \qquad
\mean{A_{ud}}=\Lambda_f^4\left(P_0+P_{int}\right)\,,
\end{array}
\label{Invariant3F_Fund}
\end{equation}
where $P_0$ and $P_{int}$ encode the angular dependence,
\begin{eqnarray}
&&P_0\equiv-\sum_{i<j} \left(y_{u_i}^2-y_{u_j}^2 \right) \left( y_{d_i}^2-y_{d_j}^2 \right) \sin^2 \theta_{ij}\,,
\label{P0}\\
&&\begin{split}
P_{int}\equiv&
\sum_{i<j,k}\left(y_{d_i}^2-y_{d_k}^2\right)\left(y_{u_j}^2-y_{u_k}^2\right)\sin^2 \theta_{ik}\sin^2\theta_{jk} \, +\\
&-\left( y_d^2-y_s^2 \right)\left( y_c^2-y_t^2 \right) \sin^2\theta_{12} \sin^2\theta_{13} \sin^2\theta_{23} \,+ \\[3mm]
&+\dfrac{1}{2}\left(y_d^2-y_s^2\right)\left(y_c^2-y_t^2\right)\cos\delta\,\sin2\theta_{12}\sin2\theta_{23}\sin\theta_{13}\,,
\end{split}
\label{Pint}
\end{eqnarray}
with $i,j,k=1,2,3$. $P_0$ generalizes the expression found in the two-family case - see eq.~(\ref{InvariantsExplicit2F_BiFund}) - containing all the terms
with a single angular dependence. The second piece, instead, $P_{int}$, contains all contributions that involve more
than one mixing angle. Notice that in this case the Jarlskog invariant appears only at the non-renormalizable level.
The most general scalar potential at the renormalizable level is now given by
\begin{equation}
V^{(4)}=\sum_{i=u,d}\left(-\mu_i^2A_i+\tilde{\mu}_iB_i+\lambda_iA_i^2+\lambda'_iA_{ii} \right)+g_{ud}A_uA_d +
\lambda_{ud}A_{ud}\,.
\label{newV4Bifund}
\end{equation}
Notice that the invariants $B_{u,d}$ have mass dimension three (instead of two for the two-generation case), so that no $B^2_{u,d}$ term can be introduced at this level.
The solutions that minimize this scalar potential have a pattern very similar
to that for the two-family case: i) no mixing is favored~\footnote{However, due to the peculiar structure of the last term in eq.~(\ref{Pint}), minima with
non-vanishing angles are now allowed, although leading to solutions which are both fine-tuned and overall physically incorrect.} , ii) all eigenvalues
tend to be degenerate in most of the parameter space. Now however, there is
a region in parameter space for which a hierachical solution is allowed for non
strictly zero, but constrained, $\tilde \mu$ . This solution has one nonvanishng Yukawa
eigenvalue per up and down sector, but to recover the hierarchy among top
and bottom masses it is necessary to further demand $g_{ud} < y^2_b /y_t^2$ which, in the absence of {\it ad hoc} symmetries, results in a
similar degree of �ne-tunnig to that for the two-family case. Furthermore, alike
to the case of an initial vanishing $\sin \theta$ at the renormalizable level for two
families, it cannot be corrected by non-renormalizeble
terms in the potential.
As in sect. \ref{sec:Bi_Fund2F} for two generation, we studied the contributions of non-renormalizable operators in the
scalar potential, with similar conclusion: the introduction of higher order terms does not lead to a more natural
description of the physical parameters. Nevertheless, some improvement can be obtained when discussing the scenario
with a fine-tuned choice of the parameters $g_{ud}$ and $\tilde{\mu}$. In this case, in fact, lighter Yukawas can
be introduced through higher order operators, even if no natural hierarchy between the first two families can be obtained.
In summary, for three generations, to consider bi-fundamental scalars (as in the case of $d=5$ Yukawa operator) alone as
the possible dynamical origin of Yukawa couplings does not lead naturally to a satisfactory pattern of masses and mixings~\footnote{See note added in proof}.
\mathversion{bold}
\subsection{$d=6$ Yukawa operator: the fundamental approach}
\label{sec:Fund3F}
\mathversion{normal}
We deal now with the case of flavons transforming in the fundamental of the flavour group $G_f$. For most of the
conventions we refer to the two-family treatment done in sect.~\ref{sec:Fund2F}.
To account for non-trivial mixing, it is necessary to introduce at least four flavons, corresponding to up and down,
left and right flavons:
\begin{equation}
\chi_u^L \in \left( 3, 1, 1\right)\,,\qquad\quad
\chi_u^R \in \left(1, 3, 1\right)\,,\qquad\quad
\chi_d^L \in \left( 3, 1, 1\right)\,,\qquad\quad
\chi_d^R \in \left(1, 1, 3\right)\,.
\label{listfundamentals3d}
\end{equation}
When they develop vevs, the flavour symmetry is spontaneously broken and the Yukawa matrices
are given as in eq.~(\ref{YukawasVEVs_Bi_Fund_2F}).
Without loss of generality, it is possible to write:
\begin{equation}
\mean{\chi_{u,d}^{L,R}} \equiv\left|\chi_{u,d}^{L,R}\right|\mathcal{V}_{L,R}^{(u,d)}
\left(\begin{array}{c}
0\\
0\\
1
\end{array}\right)\,,
\label{VEV3F_Fund}
\end{equation}
where $\mathcal{V}_{L,R}^{(u,d)}$ are $3\times3$ unitary matrices. Similarly to what was shown in sect.~\ref{sec:Fund2F},
removing the unphysical parameters, the following expressions for the Yukawa matrices are obtained:
\begin{equation}
Y_D=\frac{\left|\chi_d^L\right|\left|\chi_d^R\right|}{\Lambda_f^2} \left(\begin{array}{ccc}
0 & 0 & 0\\
0 & 0 & 0\\
0 & 0 & 1\\
\end{array}\right)\,,\qquad\quad
Y_U=\frac{\left|\chi_u^L\right|\left|\chi_u^R\right|} {\Lambda_f^2} \mathcal{V}_L^{(d)\dag} \mathcal{V}_L^{(u)}
\left(\begin{array}{ccc}
0 & 0 & 0\\
0 & 0 & 0\\
0 & 0 & 1\\
\end{array}\right)\,.
\label{Yukawas3F_Fund}
\end{equation}
This illustrates that, independently of the parametrization chosen, $Y_D$ and $Y_U$ can have only one non-vanishing
eigenvalue, as they result from multiplying two vectors. For obvious reasons, in eq.~(\ref{Yukawas3F_Fund}) the massive
state is chosen to be that of the third generation.
The flavon vevs have not broken completely the flavour symmetry, leaving a residual $SU(2)_{Q_L}\times SU(2)_{D_R}\times
SU(2)_{U_R}$ symmetry group. As a consequence any rotation in the $12$ sector is unphysical and the only physical
angle, given by the misalignment between $\mean{\chi_u^L}$ and $\mean{\chi_d^L}$ in the flavour space, can be identified
with the $23$ CKM mixing angle:
\begin{equation}
\mathcal{V}_L^{(d)\dag} \mathcal{V}_L^{(u)}=\left(
\begin{array}{ccc}
1&0&0\\
0&\cos{\theta_{23}}&\sin{\theta_{23}}\\
0&-\sin{\theta_{23}}&\cos{\theta_{23}}\\
\end{array}\right)\,.
\end{equation}
The analysis of the scalar
potential follows exactly that in sect.~\ref{sec:Fund2F} for two families (see for example eq.~(\ref{YukawasResults})),
with the obvious replacement of $y_c$, $y_s$ for $y_t$, $y_b$ and with the physical mixing
angle corresponding now to $\theta_{23}$. Both the largest hierarchy and a $\cos \theta_{23}$ naturally of ${\mathcal O}(1)$
are beautifully explained without any fine-tuning. However, as in the two-family case, it is not possible to generate
lighter fermion masses either introducing non-renormalizable interactions or adding extra RH flavons.
Nevertheless, the partial breaking of flavour symmetry provided by eq.~({\ref{Yukawas3F_Fund}}) can open
quite interesting possibilities from a model-building point of view. Consider as an example the following
multi-step approach. In a first step, only the minimal number of fundamental fields are introduced: i.e. $\chi^L$, $\chi^R_u$
and $\chi^R_d$. Their vevs break $G_f=SU(3)^3$ down to $SU(2)^3$, originating non-vanishing Yukawa couplings
only for the top and the bottom quarks, without any mixing angle (as we have only one left-handed flavon). As a second step,
four new $G_f$-triplet fields $\chi^{\prime L,R}_{u,d}$ are added, whose contributions to the Yukawa terms are suppressed
relatively to the previous flavons (i.e. $\mean \chi' \ll \mean \chi$). If their vevs
point in the direction of the unbroken flavour subgroup
$SU(2)^3$, then the residual symmetry is further reduced. As a result, non-vanishing charm and strange Yukawa couplings
are generated together with a mixing among the first two generations:
\begin{equation}
\begin{aligned}
&Y_u\equiv \frac{\mean{\chi^L}\,\mean{\chi_u^{R\dag}}}{\Lambda_f^2}+ \frac{\mean{\chi_u^{\prime L}}\,
\mean{\chi_u^{\prime R\dag}}}{\Lambda_f^2}=\left(\begin{array}{ccc}
0&\sin{\theta}\,y_c&0\\
0&\cos{\theta}\,y_c&0\\
0&0&y_t\\
\end{array}\right)\,,\\
&Y_d\equiv \frac{\mean{\chi^L}\,\mean{\chi_d^{R\dag}}}{\Lambda_f^2}+ \frac{\mean{\chi_d^{\prime L}}\,
\mean{\chi_d^{\prime R\dag}}}{\Lambda_f^2}=\left(\begin{array}{ccc}
0&0&0\\
0&y_s&0\\
0&0&y_b\\
\end{array}\right)\,.
\end{aligned}
\end{equation}
The relative suppression of the two sets of flavon vevs correspond to the hierarchy between $y_c$ and $y_t$ ($y_s$ and $y_b$)\footnote{Alternatively, all flavon vevs of similar magnitude with different flavour scale would lead to the same pattern.}. Hopefully, a
refinement of this argument would allow to explain the rest of the Yukawas and the remaining angles. The
construction of the scalar potential for such a setup would be quite model dependent though, and beyond the scope of this paper.
\subsection{Combining fundamentals and bi-fundamentals}
\label{sec:Bi_Fund-Fund}
Until now we have considered separately Yukawa operators of dimension $d=5$ and $d=6$. It is, however, interesting
to explore if some added value from the simultaneous presence of both kinds of operators can be obtained.
This is a sensible choice from the point of view of effective Lagrangians in which, working at ${\mathcal O}(1/\Lambda_f^2)$, contributions of four types may be included: i) the leading $d=5$ ${\mathcal O}(1/\Lambda_f)$ operators; ii) renormalizable terms stemming from fundamentals (i.e. from $d=6$ ${\mathcal O}(1/\Lambda_f^2)$ operators); iii) ${\mathcal O}(1/\Lambda_f^2)$ of the form $\Sigma_{u,d} \Sigma_R$ if $\Sigma_R$ turns out to be present in the spectrum; iv) other corrections numerically competitive at the orders considered here. We focus here as illustration on the impact of i) and ii):
\begin{equation}
\mathscr{L}_Y=\ov{Q}_L\left[\dfrac{\Sigma_d}{\Lambda_f}+
\dfrac{\chi_d^L\chi_d^{R\dagger} }{\Lambda_f^2} \right]D_RH +
\ov{Q}_L\left[\dfrac{\Sigma_u}{\Lambda_f}+
\dfrac{\chi_u^L \chi_u^{R\dagger}}{\Lambda_f^2} \right]U_R\tilde{H}+\text{h.c.}\,,
\label{LagrangianMixed}
\end{equation}
As the bi-fundamental flavons arise at first order in the $1/\Lambda_f$ expansion, it is suggestive to think of the fundamental
contributions as a ``higher order'' correction.
Let us then consider the case in which
the flavons develop vevs as follows:
\begin{equation}
\dfrac{\mean{\Sigma_{u,d}}}{\Lambda_f}\sim\left(
\begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
0 & 0 & y_{t,b} \\
\end{array}
\right)\,,\qquad\qquad
\dfrac{\mean{\chi_{u,d}^L}}{\Lambda_f^2}\sim \left(
\begin{array}{ccc}
0 \\
y_{c,s} \\
0 \\
\end{array}
\right)\,,
\label{lastequation}
\end{equation}
and $\chi_{u,d}^R$ acquire arbitrary vev values, although ${\mathcal O}(1)$, for all components. Nevertheless, it is important
to recall that the bi-fundamentals $\Sigma$ point in most cases to degenerate Yukawa eigenvalues instead of the pattern
in the left-hand side of eq.~(\ref{lastequation}), and either restrictive conditions on the parameters, or an extra
symmetry, have to be imposed to obtain it, see sects. 2.1.1 and 3.2. Finally,
\begin{equation}
\begin{aligned}
&Y_u=\left(\begin{array}{ccc}
0&\sin{\theta_c}\,y_c&0\\
0&\cos{\theta_c}\,y_c&0\\
0&0&y_t\\
\end{array}\right)\,,\qquad
&Y_d
\left(\begin{array}{ccc}
0&0&0\\
0&y_s&0\\
0&0&y_b\\
\end{array}\right)\,.
\end{aligned}
\end{equation}
This seems an appealing pattern, with masses for the two heavier generations and one sizable mixing angle, that we chose to identify here with the Cabibbo angle\footnote{Similar constructions have been suggested also in other contexts as in \cite{Berezhiani:2005tp,FKR:AnarchicalMasses,CFRZ:UnifiedAnarchicalMasses}.}. As for the lighter family, non-vanishing masses for the up and down quarks could now result from non-renormalizable operators.
The drawback of these combined analysis is that the direct connection between the minima of the potential and the spectrum is lost and the analysis of the potential would be very involved.
\section{Conclusions}
\label{sec:concl}
The ansatz of MFV implicitly assumes a dynamical origin for the SM Yukawa couplings. In this paper we explored
such a possibility. The simplest dynamical realization of MFV
is to identify the Yukawa couplings with the vevs of some dynamical fields, the flavons. For instance,
the Yukawa interactions themselves could result, after spontaneous symmetry breaking, from effective operators of
dimension $d>4$ invariant under the flavour symmetry, which involve one or more flavons together with the usual SM fields.
Only a scalar field (or an aggregate of fields in a scalar configuration) can get a vev, which should correspond
to the minimum of a potential. What may be the scalar potential of the MFV flavons? May some of its minima naturally correspond to the SM spectra of masses and mixing angles? These are the questions addressed in this work.
First of all, we showed here that the underlying flavour symmetry - under which the terms in the potential have to be
invariant - is a very restrictive constraint: at the renormalizable level only a few terms are allowed in the potential,
and even at the non-renormalizable level quite constrained patterns have to be respected.
The simplest realization is obtained by a one-to-one correspondence of each Yukawa coupling with a single scalar
field transforming
in the bi-fundamental of the flavour group. In the language of effective Lagrangians, this may correspond to the
lowest order terms in the flavour expansion: $d=5$ effective Yukawa operators made out of one flavon field plus
the usual SM fields. We have constructed the general scalar potential for bi-fundamental flavons, both for the case of
two and three families. At the renormalizable level, at the minimum of the potential only vanishing or undetermined
mixing angles are allowed. The introduction of either additional {\it ad hoc} symmetries or the restriction to a contrived region of the parameter domain could allow
to obtain solutions with vanishing Yukawa couplings for all quarks but those in the heaviest family. Still, mixing would
be absent. The addition of non-renormalizable terms to the potential would allow masses for the lighter families, although
without providing naturally a correct pattern of masses and mixings.
In resume, the sole consideration of flavons in the bi-fundamental representation of the flavour group does not naturally
lead to a satisfactory dynamical description of the SM quark flavour sector, at least at the classical level.
Another avenue explored in this work associates two vector flavons to each Yukawa spurion, i.e. a Yukawa
$Y\sim \mean{\chi^L}\mean{\chi^{R\dag}}/\Lambda_f^2$. This is a very attractive scenario in that while
Yukawas are composite objects, the new fields are in the fundamental representation of the flavour group,
in analogy with the case of quarks. Those flavons could be scalars or fermions: we focused exclusively on scalars.
From the point of view of effective Lagrangians, this case could correspond to the next-to leading order term
in the expansion: $d=6$ Yukawa operators. We have constructed the general scalar potential for scalar flavons
in the fundamental representation, both for the case of two and three families of quarks. By construction, this scenario
results unavoidably in a strong hierarchy of masses: at the renormalizable level only one quark gets mass in each quark
sector: they could be associated with the top and bottom quark for instance. Non-trivial mixing requires as expected a
misalignment between the flavons associated to the up and down left-handed quarks. In consequence, the minimal field
content corresponds to four fields $\chi^L_u$, $\chi^L_d$, $\chi^R_u$ and $\chi^R_d$, and the physics of mixing lies
in the interplay of the first two. In resume, for fundamental flavons it follows in a completely natural way:
i) a strong mass hierarchy between quarks of the same charge, pointing to a distinctly heavier quark in each sector;
ii) one non-vanishing mixing angle, which can be identified with the Cabibbo angle in the case of two generations,
and for instance with the rotation in the $23$ sector of the CKM matrix in the case of three generations.
Nevertheless, to achieve non-vanishing Yukawa couplings for the lighter quarks and the full mixing pattern requires,
at least at the classical level explored here, more complicated scenarios and variable degrees of fine-tuning.
Interesting possibilities which we started to explore here include replicas of fundamental flavons, in several varieties.
An intriguing one consists in considering the minimal set of $\chi$ fields plus their replicas: it allows a double step symmetry breaking
mechanism, which may produce the hierarchical quark spectrum and the shell-like pattern of the CKM matrix.
Finally, we briefly explored the possibility of introducing simultaneously bi-fundamen\-tals and fundamentals flavons.
It is a very sensible possibility from the point of view of effective Lagrangians to consider both $d=5$ and
$d=6$ Yukawa operators when working to ${\mathcal O}(1/\Lambda_f^2)$. It suggests that $d=5$ operators, which bring in the
bi-fundamentals, could give the dominant contributions, while the $d=6$ operator - which brings in the fundamentals -
should provide a correction inducing the masses of the two lighter families and the Cabibbo angle. It requires, though,
to appeal to a discrete symmetry or to restrict the parameters to the potential to a contrived region to avoid quark
mass degeneracies induced by the bi-fundamental flavons.
Overall, it is remarkable that the requirement of invariance under the flavour symmetry strongly constraints
the scalar potential of MFV, up to the point that the obtention of quark mass hierarchies and mixing angles is far
from trivial. Furthermore, besides exploring the - disappointing - impact in mixing of bi-fundamental flavons,
this work has shown that flavons in the fundamental are instead a tantalizing avenue to induce hierarchies and
non-trivial fermion mixing. A long path remains ahead, though, to naturally account for the complete observed
fermion spectrum and mixings.
\section*{Acknowledgements}
We are specially indebted to Enrico Nardi for fruitful discussions and suggestions. We also thank
Alvaro de Rujula and Pilar Hern\'andez for illuminating discussions. L. Merlo and S. Rigolin thank
the Departamento de F\'isica Te\'orica of the Universidad Aut\'onoma de Madrid for hospitality during
the development of this project.
R. Alonso and M.B. Gavela acknowledge CICYT through the project FPA2009-09017 and by CAM through the project
HEPHACOS, P-ESP-00346. R. Alonso acknowledges financial support from the MICINN grant BES-2010-037869.
L. Merlo acknowledges the German `Bundesministerium f\"ur Bildung und Forschung' under contract 05H09WOE.
S. Rigolin acknowledges the partial support of an Excellence Grant of Fondazione Cariparo and of the
European Program‚ Unification in the LHC era‚ under the contract PITN- GA-2009-237920 (UNILHC).
\section*{Note added in proof}
After this work was submitted, a paper appeared in the arXiv \cite{Nardi:2011st} where it has been suggested that
the introduction of Coleman-Weinberg quantum corrections to our results for the bi-fundamental case could generate
subdominant Yukawa splittings.
The author, using a slightly modified version of our notation, re-derived our renormalizable potential for the bi-fundamental case for three families. Looking at eq.~(\ref{newV4Bifund}) and eqs.~(\ref{Invariants2F_BiFund})--(\ref{ScalarPot2F_BiFund}), it is easy to identify the relations to move from one notation to the other:
\begin{eqnarray}
&& A_u \rightarrow T_u \, , \quad B_u \rightarrow D_u \, , \quad A'_{uu} = (A_u^2 - A_{uu}) \rightarrow 2\, A_u
\nonumber \\
&& \mu_u\rightarrow m_u \, , \quad \tilde{\mu}_u\rightarrow 2\,|\tilde{\mu}_u| \, , \quad \lambda_u\rightarrow
\lambda_u+\dfrac{1}{2}\tilde{\lambda}'_u \, , \quad \lambda'_u\rightarrow -\dfrac{1}{2}\tilde{\lambda}'_u\,.
\nonumber
\end{eqnarray}
The freedom on the relative sign between the determinant and $\tilde{\mu}$ terms allowed in our paper has been retaken in v2 of \cite{Nardi:2011st} as a
cosine dependence, which now allows negative coefficients and redefines their norm.
The hierarchical solution in which that paper is based, was already identified in our work, together with the
degenerate one. To quantify the validity range of the two solutions we found, we add here a detailed analysis
of the stability of the potential, that was not included in our previous version.
\begin{description}
\item A) {\it Stability condition for two families.} \\
For the two-family case the extremality equations read:
\begin{equation}
\frac{\partial V_u}{\partial y_c}=4\,\lambda_u\,\Lambda_{fl}^2\,y_c\left(A_u-\frac{\mu^2_u}{2\,\lambda_u}\right)+
2\,\tilde{\lambda}_u\,\Lambda_{fl}^2\,y_u\left(B_u-\frac{\tilde{\mu}_u^2}{2\,\tilde{\lambda}_u}\right)-h_u\,\Lambda_{fl}^2\left(
y_u\,A_u+2\,y_c\,B_u\right)=0 \nonumber
\end{equation}
\begin{equation}
\frac{\partial V_u}{\partial y_u}=4\,\lambda_u\,\Lambda_{fl}^2\,y_u\left(A_u-\frac{\mu^2_u}{2\,\lambda_u}\right)+
2\,\tilde{\lambda}_u\,\Lambda_{fl}^2\,y_c\left(B_u-\frac{\tilde{\mu}^2_u}{2\,\tilde{\lambda}_u}\right)-h_u\,\Lambda_{fl}^2
\left(y_c\,A_u\,+\,2\,y_u\,B_u\right)=0 \, .\nonumber
\end{equation}
One can easily verify that the hierarchical pattern $(0,\,y)$ is not a solution of these equations,
unless a severe fine-tuning on the parameters $\mu_u$, $\tilde \mu_u$, $\lambda_u$ and $h_u$ is introduced.
Only the symmetric solution $(y,\,y)$ arises as a natural minimum.
\item B) {\it Stability condition for three families.} \\
The conditions defining the minima now read as follows.
\begin{enumerate}
\item The parameter region in which only the symmetric solution $(y,\,y,\,y)$ provides a stable minimum is defined by
\begin{equation}
\frac{\tilde{\mu}^2_u}{\mu^2_u}>\frac{8\,\lambda_u'^2}{\lambda_u\,+\,\lambda'_u}\,
\nonumber
\end{equation}
for $\lambda'_u < 0$. On the other hand, for $\lambda'_u > 0$, the configuration $(y,\,y,\,y)$ is a stable minimum for any value of $\tilde{\mu}^2_u/\mu^2_u$.
\item The parameter region in which the symmetric solution is the absolute minimum, while the hierarchical configuration $(0,\,0,\,y)$ is a local minimum corresponds to
\begin{equation}
8\left(\lambda_u+\lambda_u'\right)\left(\left(4-2\frac{\lambda'_u}{\lambda_u+\lambda_u'}\right)^{3/2}
-\left(8-6\frac{\lambda_u'}{\lambda_u+\lambda_u'}\right)\right) <
\frac{\tilde{\mu}^2_u}{\mu^2_u}<\frac{8\,\lambda_u'^2}{\lambda_u+\lambda'_u}
\end{equation}
\item The parameter region in which the symmetric solution is a local minimum, while the hierarchical solution is the absolute minimum is defined by
\begin{equation} \frac{8\,\lambda_u'^2}{3\lambda_u+2\lambda'_u}<\frac{\tilde{\mu}^2_u}{\mu^2_u}<
8\left(\lambda_u+\lambda_u'\right)\left(\left(4-2\frac{\lambda'_u}{\lambda_u+\lambda_u'}\right)^{3/2}
-\left(8-6\frac{\lambda_u'}{\lambda_u+\lambda_u'}\right)\right)\,.
\end{equation}
\item Finally, the parameter region in which only the hierarchical configuration is a minimum corresponds to
\begin{equation}
\frac{\tilde{\mu}^2_u}{\mu^2_u}<\frac{8\,\lambda_u'^2}{3\lambda_u+2\lambda'_u}\,.
\end{equation}
\end{enumerate}
\begin{figure}[t]
\centering
\includegraphics[width=4in]{Fig_Stability.pdf}
\caption{Parameter space for the symmetric and hierarchical configurations. The dark-Orange corresponds to a region where the symmetric
solution $(y,y,y)$ is the stable absolute minimum, while the hierarchical solution $(0,0,y)$ is a saddle point. In the light-Orange
region, the symmetric configuration is the absolute minimum, while the hierarchical solution is a local one. On the contrary, in the light-Brown
region the symmetric configuration is a local minimum, while the hierarchical solution is an absolute one. Finally, in the dark-Brown region only the hierarchical solution is a minimum. In the plot the value $\lambda_u=1/2$ has been used for illustration.}
\label{fig:note}
\end{figure}
\end{description}
As illustrated in fig.~(\ref{fig:note}) (see also \cite{Planck2011}), for a typical $\lambda_u$ value in the perturbative
regime the symmetric configuration is the absolute minimum for most of the parameter space (here shown in dark and light orange). However, even when the hierarchical solution is preferred, yielding non-zero top and bottom Yukawa eigenvalues only, the hierarchy among up and down sectors must be considered. In particular the presence of the term $g_{ud}\,A_{u}\,A_{d}$ must be constrained by setting $g_{ud}\,<\,y_b^2\,/\,y_t^2\,\sim\,10^{-3}$ to warranty the top-bottom mass hierarchy. As already stated before, such a fine-tuning can be justified through additional symmetries. In particular in \cite{Nardi:2011st}, it is placed in the vev of flavons transforming under Abelian factors.
|
\section{Introduction}
\label{S:intro}
Let $(X,\| \cdot \|)$ be a separable Banach space, $\gamma \in \mathcal{P}(X)$ be an infinite dimensional Gaussian measure and
$H(\gamma)$ be the corresponding Cameron-Martin space with Hilbertian norm $\|\cdot \|_{H(\gamma)}$. Consider two
probability measures $\mu, \nu \in \mathcal{P}(X)$.
We will prove the existence of a solution for the following Monge minimization problem
\begin{equation}\label{E:trasp}
\min_{T: T_{\sharp}\mu = \nu} \int_{X} \| x- T(x) \|_{H(\gamma)} \mu (dx),
\end{equation}
provided $\mu$\ and $\nu$ are both absolutely continuous w.r.t. $\gamma$.
Before giving an overview of the paper, we recall the main results on the Monge problem.
In the original formulation given by Monge in 1781 the problem was settled in $\mathbb{R}^{d}$,
with the cost given by the Euclidean norm and the measures $\mu, \nu$ supposed to be absolutely
continuous and supported on two disjoint compact sets.
The original problem remained unsolved for a long time.
In 1978 Sudakov \cite{sudak} claimed to have a solution for any distance cost function induced
by a norm: an essential ingredient in the proof was that if $\mu \ll \mathcal{L}^{d}$ and $\mathcal{L}^{d}$-a.e. $\mathbb{R}^{d}$ can be decomposed into
convex sets of dimension $k$, then
then the conditional probabilities are absolutely continuous with respect to the $\mathcal{H}^{k}$ measure of the correct dimension.
But it turns out that when $d>2$, $0<k<d-1$ the property claimed by Sudakov is not true. An example with $d=3$, $k=1$ can be found in \cite{larm}.
The Euclidean case has been correctly solved only during the last decade. L. C. Evans and W. Gangbo in \cite{evagangbo}
solve the problem under the assumptions that
$\textrm{spt}\,\mu \cap \textrm{spt}\,\nu = \emptyset$, $\mu,\nu \ll \mathcal{L}^{d}$ and their densities are Lipschitz functions with compact support.
The first existence results for general absolutely continuous measures $\mu,\nu$ with compact support is independently obtained by
L. Caffarelli, M. Feldman and R.J. McCann in \cite{caffafeldmc} and by N. Trudinger and X.J. Wang in \cite{trudiwang}.
M. Feldman and R.J. McCann \cite{feldcann:mani} extend the results to manifolds with geodesic cost.
The case of a general norm as cost function on $\mathbb{R}^{d}$, including also the case with non strictly convex unitary ball,
is solved first in the particular case of crystalline norm by L. Ambrosio, B. Kirchheim and A. Pratelli in
\cite{ambprat:crist}, and then in fully generality independently by L. Caravenna in \cite{caravenna:Monge} and by T. Champion and L. De Pascale in
\cite{champdepasc:Monge}. The Monge minimization problem for non-branching geodesic metric space is studied in \cite{biacava:streconv},
where the existence is proven for spaces satisfying a finite dimensional lower curvature bound.
\subsection{Overview of the paper}
\label{Ss:over}
The approach to this problem is the one of \cite{biacava:streconv}: assume that there exists a transference plan of finite cost, then we can
\begin{enumerate}
\item reduce the problem to transportation problems along distinct geodesics;
\item show that the disintegration of the marginal $\mu$ on each geodesic is continuous;
\item find a transport map on each geodesic and piece them together.
\end{enumerate}
Indeed, since the cost function is lower semi-continuous, the existence of transference plan of finite cost implies the existence of an optimal transference plan.
This permits to reduce the minimization problem to one dimensional minimization problems.
There an explicit map can be constructed provided the first marginal measure $\mu$ is continuous (i.e. without atoms), for example
choose the monotone minimizer of the quadratic cost $|\cdot|^2$. The third point is an application of selection theorems.
All this strategy has already implemented in fully generality in \cite{biacava:streconv}.
Therefore to obtain the existence of an optimal transference map we have to show that point (2) of the strategy is fulfilled for $\mu$ and $\nu$
absolute continuous with respect to the infinite dimensional Gaussian measure $\gamma$.
We recall the main steps of the reduction to geodesics.
The geodesics used by a given transference plan $\pi$ to transport mass can be obtained from a set $\Gamma$ on which $\pi$ is concentrated.
It is well-known that every optimal transference plan is concentrated on a $\| \cdot \|_{H}$-cyclically monotone set.
Since the considered norm is non-branching, $\Gamma$
yields a natural partition $R$ of a subset of the transport set $\mathcal T_e$, i.e. the set of points on the geodesics used by $\pi$:
defining
\begin{itemize}
\item the set $\mathcal T$ made of inner points of geodesics,
\item the set $a \cup b := \mathcal T_e \setminus \mathcal T$ of initial points $a$ and end points $b$,
\end{itemize}
the cyclical monotonicity of $\Gamma$ implies that the geodesics used by $\pi$ are a partition on $\mathcal T$.
In general in $a$
there are points from which more than one geodesic starts and in $b$ there are points in which more than one geodesic ends, therefore
the membership to a geodesic can't be an equivalence relation on the set $a\cup b$.
Take as example the unit circle with $\mu = \delta_0$ and $\nu = \delta_\pi$.
We note here that $\pi$ gives also a direction along each component of $R$ and w.l.o.g. we can assume that $\mu(b) = \nu(a)=0$.
Even if we have a natural partition $R$ in $\mathcal T$ and $\mu(a) = 0$, we cannot reduce the transport problem to one dimensional problems: a necessary and sufficient condition is that the disintegration of the measure $\mu$ is strongly consistent, which is equivalent to the fact that there exists a $\mu$-measurable quotient map $f : \mathcal T \to \mathcal T$ of the equivalence relation $R$. If this is the case, then
\[
m := f_\sharp \mu, \quad \mu = \int \mu_y m(dy), \quad \mu_y(f^{-1}(y)) = 1,
\]
i.e. the conditional probabilities $\mu_y$ are concentrated on the counterimages $f^{-1}(y)$ (which are single geodesics).
In our setting the strong consistency of the disintegration of $\mu$
is a consequence of the topological properties of the geodesics of $\|\cdot\|_{H(\gamma)}$ considered as curves in $(X,\|\cdot \|)$.
Finally we obtain the one dimensional problems by
partitioning $\pi$ w.r.t. the partition $R \times (X \times X)$,
\[
\pi = \int \pi_y m(dy), \quad \nu = \int \nu_y m(dy) \quad \nu_y := (P_2)_\sharp \pi_y,
\]
and considering the one dimensional problems along the geodesic $R(y)$ with marginals $\mu_y$, $\nu_y$ and cost the arc length on the geodesic.
At this point we can study the problem of the regularity of the conditional probabilities $\mu_y$.
A natural operation on sets can be considered: the evolution along the transport set.
If $A$ is a subset of $\mathcal {T}_{e}$, we denote by $T_t(A)$ the set $T_{t}(\Gamma \cap A \times X)$ where $T_{t}$ is the map from $X \times X$ to $X$
that associates to a couple of points its convex combination at time $t$.
It turns out that the fact that $\mu(a) = 0$ and the measures $\mu_y$ are continuous depends on the behave of the function $t \mapsto \gamma(T_t(A))$.
\begin{theorem}[Proposition \ref{P:puntini} and Proposition \ref{P:nonatoms}]
\label{T:-1}
If for every $A$ with $\mu(A)>0$ there exists a sequence $t_{n} \searrow 0$ and a positive constant $C$ such that
$\gamma(T_{t_{n}}(A)) \geq C \mu(A)$, then $\mu(a) = 0$ and the conditional probabilities $\mu_y$ and $\nu_{y}$ are continuous.
\end{theorem}
This result implies that the existence of a minimizer of the Monge problem is equivalent to the regularity properties of $t \mapsto \gamma(T_t(A))$.
Hence the problem is reduced to verify that the Gaussian measure $\gamma$ satisfies the assumptions of Theorem \ref{T:-1}.
Actually this is the key part of the paper.
Let $\mu = \rho_{1} \gamma$ and $\nu =\rho_{2} \gamma$ and assume that $\rho_{1}$ and $\rho_{2}$ are bounded.
Then we find suitable $d$-dimensional measures $\mu_{d}, \nu_{d}$, absolute continuous w.r.t. the $d$-dimensional Gaussian measure $\gamma_{d}$,
converging to $\mu$ and $\nu$ respectively, such that (Theorem \ref{T:stimapprox}) $\gamma_{d}$ verifies $\gamma_{d}(T_{d,t}(A)) \geq C \mu_{d}(A)$ where the evolution now is induced by the transport problem between $\mu_{d}$ and $\nu_{d}$ and the constant $C$ does not depend on the dimension.
Passing to the limit as $d \nearrow + \infty$, we prove the same property for $\gamma$.
Hence the existence result is proved for measures with bounded densities.
To obtain the existence result in fully generality we observe that the transport set $\mathcal{T}_{e}$ is a transport set
also for suitable transport problems between measures satisfying the uniformity condition stated above (Proposition \ref{P:noiniz} and Proposition \ref{P:noatom}).
The assumption that both $\mu$ and $\nu$ are a.c. with respect to $\gamma$ is fundamental.
Indeed take as example a diffuse measure $\mu$ and $\nu =\delta_{x}$, then the constant in the evolution estimate induced by the optimal transference plan
will depend on the dimension and passing to the limit we loose all the informations on the evolution.
\begin{theorem}[Theorem \ref{T:esiste}]\label{T:2}
Let $\mu,\nu \in \mathcal{P}(X)$ with $\mu,\nu \ll \gamma$.
Then there exists a solution for the Monge minimization problem \eqref{E:trasp}
\[
\min_{T: T_{\sharp}\mu = \nu} \int \| x- T(x) \|_{H(\gamma)} \mu (dx).
\]
Moreover we can find $T$ invertible.
\end{theorem}
Conditions that ensure the existence of a transference plan of finite transference cost can be found in \cite{feyustu:MKWiener}.
\subsection{Structure of the paper}
\label{Ss:struct}
The paper is organized as follows.
In Section \ref{S:preli}, we recall the basic mathematical results we use. In Section \ref{Ss:univmeas} the fundamentals of projective set theory are listed. In Section \ref{S:disintegrazione} we recall the Disintegration Theorem, using the version of \cite{biacar:cmono}. Next, the basic results of selection principles are in Section \ref{Ss:sele}, and some fundamental results in optimal transportation theory are in Section \ref{Ss:General Facts}.
In Section \ref{s:wiener} we recall
the definition of the abstract Wiener space and of the infinite dimensional Gaussian measure.
Section \ref{S:Optimal} shows, omitting the proof, the construction done in \cite{biacava:streconv} on the Monge problem in a generalized non-branching
geodesic space $(X,d,d_{L})$ where $d_{L}$ is the distance cost.
Using only the $d_{L}$-cyclical monotonicity of a set $\Gamma$ we can obtain a partial order relation $G \subset X \times X$ as follows: $xGy$ iff there exists $(w,z) \in \Gamma$ and a geodesic $\gamma: [0,1] \to X$, with
$\gamma(0)=w$, $\gamma(1)=z$, such that $x$, $y$ belongs to $\gamma$ and $\gamma^{-1}(x) \leq \gamma^{-1}(y)$.
This set $G$ is analytic, and allows to define
\begin{itemize}
\item the transport ray set $R$ \eqref{E:Rray},
\item the transport sets $\mathcal T_e$, $\mathcal T$ (with and without and points) \eqref{E:TR0},
\item the set of initial points $a$ and final points $b$ \eqref{E:endpoint0}.
\end{itemize}
Moreover we show that $R \llcorner_{\mathcal T \times \mathcal T}$ is an equivalence relation, we can assume that the set of final points $b$ can be taken $\mu$-negligible. Since all these results are proved in \cite{biacava:streconv}, here we present just a schematic summary of the construction.
In Section \ref{Ss:Wiener} we show that the Wiener space fits into the general setting and that more regularity can be obtained.
Next, Section \ref{Ss:partition} recalls that the disintegration induced by $R$ on $\mathcal T$ is strongly consistent.
Using this fact, we can define an order preserving map $g$ which maps our transport problem into a transport problem on $\mathcal S \times \mathbb{R}$, where $\mathcal S$ is a cross section of $R$ (Proposition \ref{P:gammaclass}). Finally we show that under this assumption there exists a transference plan with the same cost of $\pi$ which leaves the common mass $\mu \wedge \nu$ at the same place (note that in general this operation lowers the transference cost).
In Section \ref{S:disin} we prove Theorem \ref{T:-1}.
Let $T_{t}(x,y) = x(1-t) + yt$, then we define $T_{t}(A)$ as the set $T_{t}(\Gamma \cap A \times X)$ where $\Gamma$ is the $\| \cdot \|_{H(\gamma)}$-cyclically monotone
set where the transference plan $\pi$ is concentrated.
We show that under the condition
\begin{equation}\label{E:evol}
\mu(A) > 0 \quad \Longrightarrow \quad \gamma(T_{t}(A)) \geq C \gamma(A \cap \{\rho_{1}>0 \})
\end{equation}
the set of initial points $a$ is $\mu$-negligible (Proposition \ref{P:puntini}). Then, under the same assumption, we prove that the conditional probabilities $\mu_y$
and $\nu_{y}$ are continuous (Proposition \ref{P:nonatoms}).
In Section \ref{S:approx} we prove that, choosing $\mu_{d} : = P_{d\,\sharp} \mu$ and $\nu_{d}:= P_{d\,\sharp}\nu$, if
condition \eqref{E:evol} is verified by $\gamma_{d}$ w.r.t. the evolution induced by the finite dimensional Monge problem between $\mu_{d}$ and $\nu_{d}$,
then condition \eqref{E:evol} passes to the limit provided the constant $C$ is independent on the dimension (Theorem \ref{T:approx}).
Section \ref{S:finite} proves condition \eqref{E:evol} in the finite dimensional case. It follows directly from the proof that
the constant $C$ of condition \eqref{E:evol} depends only on the lower and upper bound of the densities of the marginal measures and is independent on the
dimension.
In Section \ref{S:solu} we obtain the existence of an optimal transport map.
Proposition \ref{P:noiniz} proves that the set of initial points is $\mu$-negligible and and the set of final points is $\nu$-negligible.
Proposition \ref{P:noatom} proves that the conditional probabilities $\mu_{y}$ and $\nu_{y}$ are continuous.
Finally Theorem \ref{T:esiste} states the existence results.
We end with a list of notations, Section \ref{S:notation}.
\section{Preliminaries}
\label{S:preli}
In this section we recall some general facts about projective classes, the Disintegration Theorem for measures, measurable selection principles, geodesic spaces and optimal transportation problems.
\subsection{Borel, projective and universally measurable sets}
\label{Ss:univmeas}
The \emph{projective class $\Sigma^1_1(X)$} is the family of subsets $A$ of the Polish space $X$ for which there exists $Y$ Polish and $B \in \mathcal{B}(X \times Y)$ such that $A = P_1(B)$. The \emph{coprojective class $\Pi^1_1(X)$} is the complement in $X$ of the class $\Sigma^1_1(X)$. The class $\Sigma^1_1$ is called \emph{the class of analytic sets}, and $\Pi^1_1$ are the \emph{coanalytic sets}.
The \emph{projective class $\Sigma^1_{n+1}(X)$} is the family of subsets $A$ of the Polish space $X$ for which there exists $Y$ Polish and $B \in \Pi^1_n(X \times Y)$ such that $A = P_1(B)$. The \emph{coprojective class $\Pi^1_{n+1}(X)$} is the complement in $X$ of the class $\Sigma^1_{n+1}$.
If $\Sigma^1_n$, $\Pi^1_n$ are the projective, coprojective pointclasses, then the following holds (Chapter 4 of \cite{Sri:courseborel}):
\begin{enumerate}
\item $\Sigma^1_n$, $\Pi^1_n$ are closed under countable unions, intersections (in particular they are monotone classes);
\item $\Sigma^1_n$ is closed w.r.t. projections, $\Pi^1_n$ is closed w.r.t. coprojections;
\item if $A \in \Sigma^1_n$, then $X \setminus A \in \Pi^1_n$;
\item the \emph{ambiguous class} $\Delta^1_n = \Sigma^1_n \cap \Pi^1_n$ is a $\sigma$-algebra
and $\Sigma^1_n \cup \Pi^1_n \subset \Delta^1_{n+1}$.
\end{enumerate}
We will denote by $\mathcal{A}$ the $\sigma$-algebra generated by $\Sigma^1_1$: clearly $\mathcal{B} = \Delta^1_1 \subset \mathcal{A} \subset \Delta^1_2$.
We recall that a subset of $X$ Polish is \emph{universally measurable} if it belongs to all completed $\sigma$-algebras of all Borel measures on $X$:
it can be proved that every set in $\mathcal{A}$ is universally measurable.
We say that $f:X \to \mathbb{R} \cup \{\pm \infty\}$ is a \emph{Souslin function} if $f^{-1}(t,+\infty] \in \Sigma^{1}_{1}$.
For the proof of the following lemma see \cite{biacava:streconv}.
\begin{lemma}
\label{L:measuregP}
If $f : X \to Y$ is universally measurable, then $f^{-1}(U)$ is universally measurable if $U$ is.
\end{lemma}
\subsection{Disintegration of measures}
\label{S:disintegrazione}
Given a measurable space $(R, \mathscr{R})$ and a function $r: R \to S$, with $S$ generic set, we can endow $S$ with the \emph{push forward $\sigma$-algebra} $\mathscr{S}$ of $\mathscr{R}$:
\[
Q \in \mathscr{S} \quad \Longleftrightarrow \quad r^{-1}(Q) \in \mathscr{R},
\]
which could be also defined as the biggest $\sigma$-algebra on $S$ such that $r$ is measurable. Moreover given a measure space
$(R,\mathscr{R},\rho)$, the \emph{push forward measure} $\eta$ is then defined as $\eta := (r_{\sharp}\rho)$.
Consider a probability space $(R, \mathscr{R},\rho)$ and its push forward measure space $(S,\mathscr{S},\eta)$ induced by a map $r$. From the above definition the map $r$ is clearly measurable and inverse measure preserving.
\begin{definition}
\label{defi:dis}
A \emph{disintegration} of $\rho$ \emph{consistent with} $r$ is a map $\rho: \mathscr{R} \times S \to [0,1]$ such that
\begin{enumerate}
\item $\rho_{s}(\cdot)$ is a probability measure on $(R,\mathscr{R})$ for all $s\in S$,
\item $\rho_{\cdot}(B)$ is $\eta$-measurable for all $B \in \mathscr{R}$,
\end{enumerate}
and satisfies for all $B \in \mathscr{R}, C \in \mathscr{S}$ the consistency condition
\[
\rho\left(B \cap r^{-1}(C) \right) = \int_{C} \rho_{s}(B) \eta(ds).
\]
A disintegration is \emph{strongly consistent with respect to $r$} if for all $s$ we have $\rho_{s}(r^{-1}(s))=1$.
\end{definition}
The measures $\rho_s$ are called \emph{conditional probabilities}.
We say that a $\sigma$-algebra $\mathcal{H}$ is \emph{essentially countably generated} with respect to a measure $m$ if there exists a countably generated $\sigma$-algebra $\hat{\mathcal{H}}$ such that for all $A \in \mathcal{H}$ there exists $\hat{A} \in \hat{\mathcal{H}}$ such that $m (A \vartriangle \hat{A})=0$.
We recall the following version of the disintegration theorem that can be found on \cite{Fre:measuretheory4}, Section 452 (see \cite{biacar:cmono} for a direct proof).
\begin{theorem}[Disintegration of measures]
\label{T:disintr}
Assume that $(R,\mathscr{R},\rho)$ is a countably generated probability space, $R = \cup_{s \in S}R_{s}$ a partition of R, $r: R \to S$ the quotient map and $\left( S, \mathscr{S},\eta \right)$ the quotient measure space. Then $\mathscr{S}$ is essentially countably generated w.r.t. $\eta$ and there exists a unique disintegration $s \mapsto \rho_{s}$ in the following sense: if $\rho_{1}, \rho_{2}$ are two consistent disintegration then $\rho_{1,s}(\cdot)=\rho_{2,s}(\cdot)$ for $\eta$-a.e. $s$.
If $\left\{ S_{n}\right\}_{n\in \mathbb{N}}$ is a family essentially generating $\mathscr{S}$ define the equivalence relation:
\[
s \sim s' \iff \ \{ s \in S_{n} \iff s'\in S_{n}, \ \forall\, n \in \mathbb{N}\}.
\]
Denoting with p the quotient map associated to the above equivalence relation and with $(L,\mathscr{L}, \lambda)$ the quotient measure space, the following properties hold:
\begin{itemize}
\item $R_{l}:= \cup_{s\in p^{-1}(l)}R_{s} = (p \circ r)^{-1}(l)$ is $\rho$-measurable and $R = \cup_{l\in L}R_{l}$;
\item the disintegration $\rho = \int_{L}\rho_{l} \lambda(dl)$ satisfies $\rho_{l}(R_{l})=1$, for $\lambda$-a.e. $l$. In particular there exists a
strongly consistent disintegration w.r.t. $p \circ r$;
\item the disintegration $\rho = \int_{S}\rho_{s} \eta(ds)$ satisfies $\rho_{s}= \rho_{p(s)}$ for $\eta$-a.e. $s$.
\end{itemize}
\end{theorem}
In particular we will use the following corollary.
\begin{corollary}
\label{C:disintegration}
If $(S,\mathscr{S})=(X,\mathcal{B}(X))$ with $X$ Polish space, then the disintegration is strongly consistent.
\end{corollary}
\subsection{Selection principles}
\label{Ss:sele}
Given a multivalued function $F: X \to Y$, $X$, $Y$ metric spaces, the \emph{graph} of $F$ is the set
\begin{equation}
\label{E:graphF}
\textrm{graph}(F) := \big\{ (x,y) : y \in F(x) \big\}.
\end{equation}
The \emph{inverse image} of a set $S\subset Y$ is defined as:
\begin{equation}
\label{E:inverseF}
F^{-1}(S) := \big\{ x \in X\ :\ F(x)\cap S \neq \emptyset \big\}.
\end{equation}
For $F \subset X \times Y$, we denote also the sets
\begin{equation}
\label{E:sectionxx}
F_x := F \cap \{x\} \times Y, \quad F^y := F \cap X \times \{y\}.
\end{equation}
In particular, $F(x) = P_2(\textrm{graph}(F)_x)$, $F^{-1}(y) = P_1(\textrm{graph}(F)^y)$. We denote by $F^{-1}$ the graph of the inverse function
\begin{equation}
\label{E:F-1def}
F^{-1} := \big\{ (x,y): (y,x) \in F \big\}.
\end{equation}
We say that $F$ is \emph{$\mathcal{R}$-measurable} if $F^{-1}(B) \in \mathcal{R}$ for all $B$ open. We say that $F$ is \emph{strongly Borel measurable} if inverse images of closed sets are Borel. A multivalued function is called \emph{upper-semicontinuous} if the preimage of every closed set is closed: in particular u.s.c. maps are strongly Borel measurable.
In the following we will not distinguish between a multifunction and its graph. Note that the \emph{domain of $F$} (i.e. the set $P_1(F)$) is in general a subset of $X$. The same convention will be used for functions, in the sense that their domain may be a subset of $X$.
Given $F \subset X \times Y$, a \emph{section $u$ of $F$} is a function from $P_1(F)$ to $Y$ such that $\textrm{graph}(u) \subset F$. We recall the following selection principle, Theorem 5.5.2 of \cite{Sri:courseborel}, page 198.
\begin{theorem}
\label{T:vanneuma}
Let $X$ and $Y$ be Polish spaces, $F \subset X \times Y$ analytic, and $\mathcal{A}$ the $\sigma$-algebra generated by the analytic subsets of X. Then there is an $\mathcal{A}$-measurable section $u : P_1(F) \to Y$ of $F$.
\end{theorem}
A \emph{cross-section of the equivalence relation $E$} is a set $S \subset E$ such that the intersection of $S$ with each equivalence class is a singleton. We recall that a set $A \subset X$ is saturated for the equivalence relation $E \subset X \times X$ if $A = \cup_{x \in A} E(x)$.
The next result is taken from \cite{Sri:courseborel}, Theorem 5.2.1.
\begin{theorem}
\label{T:KRN}
Let $Y$ be a Polish space, $X$ a nonempty set, and $\mathcal{L}$ a $\sigma$-algebra of subset of $X$.
Every $\mathcal{L}$-measurable, closed value multifunction $F:X \to Y$ admits an $\mathcal{L}$-measurable section.
\end{theorem}
A standard corollary of the above selection principle is that if the disintegration is strongly consistent in a Polish space, then up to a saturated set of negligible measure there exists a Borel cross-section.
In particular, we will use the following corollary.
\begin{corollary}
\label{C:weelsupprr}
Let $F \subset X \times X$ be $\mathcal{A}$-measurable, $X$ Polish, such that $F_x$ is closed and define the equivalence relation $x \sim y \ \Leftrightarrow \ F(x) = F(y)$. Then there exists a $\mathcal{A}$-section $f : P_1(F) \to X$ such that $(x,f(x)) \in F$ and $f(x) = f(y)$ if $x \sim y$.
\end{corollary}
\begin{proof}
For all open sets $G \subset X$, consider the sets $F^{-1}(G) = P_1(F \cap X \times G) \in \mathcal{A}$, and let $\mathcal{R}$ be the $\sigma$-algebra generated by $F^{-1}(G)$. Clearly $\mathcal{R} \subset \mathcal{A}$.
If $x \sim y$, then
\[
x \in F^{-1}(G) \quad \Longleftrightarrow \quad y \in F^{-1}(G),
\]
so that each equivalence class is contained in an atom of $\mathcal{R}$, and moreover by construction $x \mapsto F(x)$ is $\mathcal{R}$-measurable.
We thus conclude by using Theorem \ref{T:KRN} that there exists an $\mathcal{R}$-measurable section $f$: this measurability condition implies that $f$ is constant on atoms, in particular on equivalence classes.
\end{proof}
\subsection{General facts about optimal transportation}
\label{Ss:General Facts}
Let $(X,\mathcal B,\mu)$ and $(Y,\mathcal B,\nu)$ be two Polish probability spaces and $c : X \times Y \to \mathbb{R}$ be a Borel measurable function. Consider the set of \emph{transference plans}
\[
\Pi (\mu,\nu) := \Big\{ \pi \in \mathcal{P}(X\times Y) : (P_1)_\sharp \pi = \mu, (P_2)_\sharp \pi = \nu \Big\}.
\]
Define the functional
\begin{equation}
\label{E:Ifunct}
\begin{array}{ccccl}
\mathcal{I} &:& \Pi(\mu,\nu) &\to& \mathbb{R}^{+} \cr
&& \pi &\mapsto& \mathcal{I}(\pi):=\int c \pi.
\end{array}
\end{equation}
The \emph{Monge-Kantorovich minimization problem} is to find the minimum of $\mathcal{I}$ over all transference plans.
If we consider a $\mu$-measurable \emph{transport map} $T : X \to Y$ such that $T_{\sharp}\mu=\nu$, the functional \eqref{E:Ifunct} becomes
\[
\mathcal I(T):= \mathcal I \big( (Id \times T)_\sharp \mu \big) = \int c(x,T(x)) \mu(dx).
\]
The minimum problem over all $T$ is called \emph{Monge minimization problem}.
The Kantorovich problem admits a (pre) dual formulation.
\begin{definition}
A map $\varphi : X \to \mathbb{R} \cup \{-\infty\} $ is said to be \emph{$c$-concave} if it is not identically $-\infty$ and there exists $\psi : Y \to \mathbb{R} \cup \{-\infty\}$, $\psi \not\equiv -\infty$, such that
\[
\varphi(x) = \inf_{y \in Y} \big\{ c(x,y) - \psi(y) \big\}.
\]
The \emph{$c$-transform} of $\varphi$ is the function
\begin{equation}
\label{E:ctransf}
\varphi^c(y) := \inf_{x\in X} \left\{ c(x,y) - \varphi (x) \right\}.
\end{equation}
The \emph{$c$-superdifferential $\partial^c \varphi$} of $\varphi$ is the subset of $X \times Y$ defined by
\begin{equation}
\label{E:csudiff}
\partial^{c}\varphi := \Big\{ (x,y) : c(x,y) - \varphi(x) \leq c(z,y) - \varphi(z) \ \forall z \in X \Big\} \subset X \times Y.
\end{equation}
\end{definition}
\begin{definition}\label{D:cicl}
A set $\Gamma \subset X \times Y$ is said to be \emph{$c$-cyclically monotone} if, for any $n \in \mathbb{N}$ and for any family $(x_0,y_0),\dots,(x_n,y_n)$ of points of $\Gamma$, the following inequality holds:
\[
\sum_{i=0}^nc(x_i,y_i) \leq \sum_{i=0}^nc(x_{i+1},y_i),
\]
where $x_{n+1} = x_0$.
A transference plan is said to be \emph{$c$-cyclically monotone} if it is concentrated on a $c$-cyclically monotone set.
\end{definition}
Consider the set
\begin{equation}
\label{E:Phicset}
\Phi_c := \Big\{ (\varphi,\psi) \in L^1(\mu) \times L^1(\nu): \varphi(x) + \psi(y) \leq c(x,y) \Big\}.
\end{equation}
Define for all $(\varphi,\psi)\in \Phi_c$ the functional
\begin{equation}
\label{E:Jfunct}
J(\varphi,\psi) := \int \varphi \mu + \int \psi \nu.
\end{equation}
The following is a well known result (see Theorem 5.10 of \cite{villa:Oldnew}).
\begin{theorem}[Kantorovich Duality]
\label{T:kanto}
Let X and Y be Polish spaces, let $\mu \in \mathcal{P}(X)$ and $\nu \in \mathcal{P}(Y)$, and let $c : X \times Y \to [0,+\infty]$ be lower semicontinuous. Then the following holds:
\begin{enumerate}
\item Kantorovich duality:
\[
\inf_{\pi \in \Pi(\mu,\nu)} \mathcal{I} (\pi) = \sup _{(\varphi,\psi)\in \Phi_{c}} J(\varphi,\psi).
\]
Moreover, the infimum on the left-hand side is attained and the right-hand side is also equal to
\[
\sup _{(\varphi,\psi)\in \Phi_{c}\cap C_{b}} J(\varphi,\psi),
\]
where $C_{b}= C_b(X, \mathbb{R}) \times C_b(Y,\mathbb{R})$.
\item If $c$ is real valued and the optimal cost is finite, then there is a measurable $c$-cyclically monotone set $\Gamma \subset X\times Y$, closed if $c$ is continuous, such that for any $\pi \in \Pi(\mu,\nu)$ the following statements are equivalent:
\begin{enumerate}
\item $\pi$ is optimal;
\item $\pi$ is $c$-cyclically monotone;
\item $\pi$ is concentrated on $\Gamma$;
\item there exists a $c$-concave function $\varphi$ such that $\pi$-a.s. $\varphi(x)+\varphi^{c}(y)=c(x,y)$.
\end{enumerate}
\item If moreover
\[
c(x,y) \leq c_{X}(x) + c_{Y}(y), \quad \ c_{X}\ \mu\textrm{-integrable}, \ c_{Y}\ \nu\textrm{-integrable},
\]
then the supremum is attained:
\[
\sup_{\Phi_c} J = J(\varphi, \varphi^{c}) = \inf_{\pi \in \Pi(\mu,\nu)} \mathcal{I}(\pi).
\]
\end{enumerate}
\end{theorem}
We recall also that if $-c$ is Souslin, then every optimal transference plan $\pi$ is concentrated on a $c$-cyclically monotone set \cite{biacar:cmono}.
\subsection{Approximate differentiability of transport maps}
The following results are taken from \cite{ambgiglsav:gradient} where are presented in fully generality.
\begin{definition}[Approximate limit and approximate differential]
Let $\Omega \subset \mathbb{R}^{d}$ be an open set and $f:\Omega \to \mathbb{R}^{m}$. We say that $f$ has an approximate limit (respectively, approximate differential) at $x \in \Omega$
if there exists a function $g:\Omega \to \mathbb{R}^{m}$ continuous (resp. differentiable) at $x$ such that the set $\{ f \neq g\}$ has Lebesgue-density 0 at $x$.
In this case the approximate limit (resp. approximate differential) will be denoted by $\tilde f (x)$ (resp. $\tilde \nabla f(x)$).
\end{definition}
Recall that if $f:\Omega \to \mathbb{R}^{m}$ is $\mathcal{L}^{d}$-measurable, then it has approximate limit $\tilde f (x)$ at $\mathcal{L}^{d}$-a.e. $x \in \Omega$
and $f(x) = \tilde f (x)$ $\mathcal{L}^{d}$-a.e..
Consider $m=d$ and denote with $\Sigma_{f}$ the Borel set of points where $f$ is approximate differential.
\begin{lemma}[Density of the push-forward]\label{L:density}
Let $\rho \in L^{1}(\mathbb{R}^{d})$ be a nonnegative function and assume that there exists a Borel set $\Sigma \subset \Sigma_{f}$
such that $\tilde f \llcorner_{\Sigma}$ is injective and $\{ \rho >0 \} \setminus \Sigma$ is $\mathcal{L}^{d}$-negligible.
Then $f_{\sharp} \rho \mathcal{L}^{d} \ll \mathcal{L}^{d}$ if and only if $|\det \tilde \nabla f|>0$ for $\mathcal{L}^{d}$-a.e. on $\Sigma$ and
in this case
\begin{equation}\label{E:area}
f_{\sharp} (\rho \mathcal{L}^{d}) = \frac{\rho}{|\det \tilde \nabla f|} \circ \tilde f^{-1}\llcorner_{f(\Sigma)} \mathcal{L}^{d}.
\end{equation}
\end{lemma}
To conclude we include a regularity result for the Monge minimization problem in $\mathbb{R}^{d}$ with cost $c_{p}(x,y) = |x-y|^{p}$, $p>1$
(Theorem 6.2.7 of \cite{ambgiglsav:gradient}):
\begin{equation}\label{E:monge}
\min_{T: T_{\sharp}\mu = \nu} \int_{\mathbb{R}^{d}} c_{p}(x,T(x)) \mu(dx).
\end{equation}
\begin{theorem}\label{T:different}
Assume that $\mu \in \mathcal{P}^{r}(\mathbb{R}^{d})$, $\nu \in \mathcal{P}(\mathbb{R}^{d})$ and
\[
\mu\bigg( \Big\{ x\in \mathbb{R}^{d} : \int c_{p}(x,y) \nu (dy)< +\infty \Big\}\bigg),
\nu\Big( \Big\{ y\in \mathbb{R}^{d} : \int c_{p}(x,y) \mu (dx) < +\infty \Big\}\bigg) >0.
\]
If the minimum of \eqref{E:Ifunct} is finite, then
\begin{itemize}
\item[$i)$] there exists a unique solution $T_{p}$ for the Monge problem \eqref{E:monge};
\item[$ii)$] for $\mu$-a.e. $x \in \mathbb{R}^{d}$ the map $T_{p}$ is approximately differentiable at $x$ and $\tilde \nabla T_{p}(x)$
is diagonalizable with nonnegative eigenvalues.
\end{itemize}
\end{theorem}
\section{The Abstract Wiener space}
\label{s:wiener}
In this section we describe our setting. The main reference is \cite{boga:gauss}.
Given an infinite dimensional separable Banach space $X$,
we denote by $\|\cdot\|_{X}$ its norm and $X^{*}$ denotes the topological dual, with duality $\langle\cdot, \cdot\rangle$.
Given the elements $x^{*}_{1}, \dots, x^{*}_{m}$ in $X^{*}$, we denote by $\Pi_{x^{*}_{1}, \dots, x^{*}_{m}} : X \to \mathbb{R}^{m}$ the map
\[
\Pi_{x^{*}_{1}, \dots, x^{*}_{m}}( x) : = \left( \langle x, x^{*}_{1}\rangle, . . . , \langle x, x^{*}_{m}\rangle\right).
\]
Denoted with $\mathcal{E}(X)$ the $\sigma$-algebra generated by $X^{*}$.
A set $C \in \mathcal{E}(X)$ is called a \emph{cylindrical set} and if
\[
C = \{ x \in X : \Pi_{\{x^{*}_{i}\}} \in B \}, \quad B \subset \mathbb{R}^{\infty}, \quad \{x^{*}_{i}\}_{i\in \mathbb{N}} \subset X^{*}
\]
we will denote the cylindrical sets with $C(B)$, and $B$ is the base of $C$. In our setting $\mathcal{B}(X)= \mathcal{E}(X)$.
Let $\gamma$ be a non-degenerate centred Gaussian measure defined on $X$. This means that $\gamma \in \mathcal{P}(X)$, is not concentrated on
a proper subspace of $X$ and for every
$x^{*} \in X^{*}$ the measure $x^{*}_{\sharp} \gamma$ is a centred Gaussian measure on $\mathbb{R}$, that is, the Fourier transform of $\gamma$
is given by
\[
\hat \gamma (x^{*}) = \int_{X} \exp\{ i \langle x^{*}, x \rangle \} \gamma(dx) = \exp \Big\{ -\frac{1}{2} \langle x^{*}, Q x^{*} \rangle \Big\}
\]
where $Q \in L(X^{*},X)$ is the covariance operator. The non-degeneracy hypothesis of $\gamma$ is equivalent to $\langle x^{*},Qx^{*} \rangle > 0$
for every $x^{*} \neq 0$. The covariance operator $Q$ is symmetric, positive and uniquely determined by the relation
\[
\langle y^{*}, Q x^{*} \rangle = \int_{X} \langle x^{*},x \rangle \langle y^{*} , x \rangle \gamma(dx), \quad \forall x^{*},y^{*}\in X^{*}.
\]
The fact that $Q$ is bounded follows from the Fernique's Theorem, see \cite{boga:gauss}. This imply that any $x^{*} \in X^{*}$ defines a
function $x \mapsto x^{*}(x)$ that belongs to $L^{p}(X,\gamma)$ for all $p\geq 1$. In particular let us denote by
$R^{*}_{\gamma} : X^{*} \to L^{2}(X,\gamma)$ the embedding $R^{*}_{\gamma}x^{*}(x): = \langle x^{*}, x \rangle$.
The space $\mathscr{H}$ given by the closure of $R^{*}_{\gamma}X^{*}$ in $L^{2}(X,\gamma)$ is called the \emph{reproducing kernel} of the Gaussian measure.
The definition is motivated by the fact that if we consider the operator $R_{\gamma} : \mathscr{H} \to X$ whose adjoint is $R_{\gamma}^{*}$ then
$Q = R_{\gamma}R_{\gamma}^{*}$:
\[
\langle y^{*}, R_{\gamma}R_{\gamma}^{*}x^{*} \rangle = \langle R_{\gamma}^{*}y^{*},R_{\gamma}^{*}x^{*} \rangle_{\mathscr{H}} =
\int_{X} \langle x^{*},x \rangle \langle y^{*} , x \rangle \gamma(dx) = \langle y^{*}, Q x^{*} \rangle.
\]
It can proven that $R_{\gamma}$ is injective, compact and
\begin{equation}\label{E:coord}
R_{\gamma}\hat h = \int_{X} \hat h(x) x \gamma(dx), \quad \hat h \in \mathscr{H},
\end{equation}
where the integral is understood in the Bochner or Pettis sense.
The space $H(\gamma) = R_{\gamma} \mathscr{H} \subset X$ is called the Cameron-Martin space. It is a separable Hilbert space with inner product inherited
from $L^{2}(X,\gamma)$ via $R_{\gamma}$:
\[
\langle h_{1} , h_{2} \rangle_{H(\gamma)} = \langle \hat h_{1} , \hat h_{2} \rangle_{\mathscr{H}}.
\]
for all $h_{1},h_{2} \in H$ with $h_{i} = R_{\gamma} \hat h_{i}$ for $i = 1,2$. Moreover $H$ is a dense subspace of $X$ and by the compactness of $R_{\gamma}$ follows that the embedding of $(H(\gamma),\| \cdot \|_{H(\gamma)})$ into $(X,\| \cdot \|)$ is compact. Note that if $X$ is infinite dimensional then $\gamma(H)=0$
and if $X$ is finite dimensional then $X=H(\gamma)$.
\subsection{Finite dimensional approximations}\label{Ss:approx}
Using the embedding of $X^{*}$ in $L^{2}(X,\gamma)$ we say that a family $\{ x^{*}_{i} \} \subset X^{*}$ is orthonormal if the corresponding
family $\{ R_{\gamma}^{*} x^{*}_{i}\}$ is orthonormal in $\mathscr{H}$.
In particular starting from a sequence $\{y^{*}_{i} \}_{i\in \mathbb{N}} $ whose image under $R_{\gamma}^{*}$ is dense
in $\mathscr{H}$, we can obtain an orthonormal basis $R_{\gamma}^{*}x^{*}_{i}$ of $\mathscr{H}$. Therefore also
$h_{j} = R_{\gamma}R_{\gamma}^{*}x^{*}_{j}$ provide an orthonormal basis in $H(\gamma)$.
In the following we will consider a fixed orthonormal basis $\{e_{i}\}$ of $H(\gamma)$ with $e_{i} = R_{\gamma} \hat e_{i}$ for $\hat e_{i} \in R_{\gamma}^{*}X^{*}$.
\begin{proposition}[\cite{boga:gauss}, Proposition 3.8.12]\label{P:appro}
Let $\gamma$ be a centred Gaussian measure on a Banach space $X$ and $\{e_{i}\}$ an orthonormal basis in $H(\gamma)$.
Define $P_{d} x : = \sum_{i=1}^{d} \langle \hat e_{i}, x\rangle e_{i}$. Then the sequence of measures $\gamma_{d}:= P_{d\,\sharp}\gamma$
converges weakly to $\gamma$.
\end{proposition}
The measure $\gamma_{d}$ defined above is a centred non-degenerate $d$-dimensional Gaussian measure
and, due to the orthonormality of $\{e_{i}\}_{i \in \mathbb{N}}$, with identity covariance matrix.
Note that from \eqref{E:coord} it follows that $\langle \hat e_{j} , x \rangle = \langle e_{i} ,x \rangle_{H}$ for all $x \in H$. Hence
we will not specify whether the measures $\gamma_{d}$ is probability measures on $\mathbb{R}^{d}$ or on $P_{d}H$:
\[
\gamma_{d} = \hat e_{1\,\sharp} \gamma \otimes \dots \otimes \hat e_{d\,\sharp} \gamma, \qquad
\hat e_{j\,\sharp} \gamma = \frac{1}{\sqrt{2\pi}} \exp\Big\{ - \frac{x^{2}}{2} \Big\} \mathcal{L}^{1}.
\]
For every $d \in \mathbb{N}$ we can disintegrate $\gamma$ w.r.t. the partition induced by the saturated sets of $P_{d}$:
\begin{equation}\label{E:disintg}
\gamma = \int \gamma_{y,d}^{\perp} \gamma_{d}(dy), \quad \gamma_{y,d}^{\perp}(P_{d}^{-1}(y))=1 \quad \textrm{for}\ \gamma_{d}-\textrm{a.e.}\ y.
\end{equation}
\section{Optimal transportation in geodesic spaces}
\label{S:Optimal}
In what follows $(X,d,d_{L})$ is a generalized non-branching geodesic space in the sense of \cite{biacava:streconv}.
In this Section we retrace, omitting the proof, the construction done in \cite{biacava:streconv} that permits to reduce the Monge problem
with non-branching geodesic distance cost $d_{L}$, to a family of one dimensional transportation problems.
The triple $(X,\| \cdot \|,\| \cdot \|_{H(\gamma)})$ is a generalized non-branching geodesic space in the sense of \cite{biacava:streconv}.
Using only the $d_L$-cyclical monotonicity of $\Gamma$, we obtain a partial order relation $G \subset X \times X$.
The set $G$ is analytic, and allows to define the transport ray set $R$, the transport sets $\mathcal T_e$, $\mathcal T$,
and the set of initial points $a$ and final points $b$.
Moreover we show that $R \llcorner_{\mathcal T \times \mathcal T}$ is an equivalence relation and that we can assume the set of final points $b$ to be $\mu$-negligible.
Let $\mu, \nu \in \mathcal{P}(X)$ and let $\pi \in \Pi(\mu,\nu)$ be a $d_L$-cyclically monotone transference plan with finite cost.
By inner regularity, we can assume that the optimal transference plan is concentrated on a $\sigma$-compact $d_L$-cyclically monotone set
$\Gamma \subset \{d_L(x,y) < +\infty\}$. By Lusin Theorem, we can require also that $d_L \llcorner_{\Gamma}$ is $\sigma$-continuous:
\begin{equation}\label{E:gamman}
\Gamma = \cup_n \Gamma_n, \ \Gamma_n \subset \Gamma_{n+1} \ \text{compact}, \quad d_L \llcorner_{\Gamma_n} \ \text{continuous.}
\end{equation}
Consider the set
\begin{align}
\label{E:gGamma}
\Gamma' :=&~ \bigg\{ (x,y) : \exists I \in \mathbb{N}_0, (w_i,z_i) \in \Gamma \ \text{for} \ i = 0,\dots,I, \ z_I = y \crcr
&~ \qquad \qquad w_{I+1} = w_0 = x, \ \sum_{i=0}^I d_L(w_{i+1},z_i) - d_L(w_i,z_i) = 0 \bigg\}.
\end{align}
In other words, we concatenate points $(x,z), (w,y) \in \Gamma$ if they are initial and final point of a cycle with total cost $0$.
One can prove that $\Gamma \subset \Gamma' \subset \{d_L(x,y) < +\infty\}$, if $\Gamma$ is analytic so is $\Gamma'$
and if $\Gamma$ is $d_L$-cyclically monotone so is $\Gamma'$.
\begin{definition}[Transport rays]
\label{D:Gray}
Define the \emph{set of oriented transport rays}
\begin{equation}
\label{E:trG}
G := \Big\{ (x,y): \exists (w,z) \in \Gamma', d_L(w,x) + d_L(x,y) + d_L(y,z) = d_L(w,z) \Big\}.
\end{equation}
For $x \in X$, the \emph{outgoing transport rays from $x$} is the set $G(x)$ and the \emph{incoming transport rays in $x$} is the set $G^{-1}(x)$. Define the \emph{set of transport rays} as the set
\begin{equation}
\label{E:Rray}
R := G \cup G^{-1}.
\end{equation}
\end{definition}
It is fairly easy to prove that $G$ is still $d_{L}$-cyclically monotone, $\Gamma' \subset G \subset \{d_L(x,y) < +\infty\}$ and
$G$ and $R$ are analytic sets.
\begin{definition} Define the \emph{transport sets}
\begin{subequations}
\label{E:TR0}
\begin{align}
\label{E:TR}
\mathcal T :=&~ P_1 \big( \textrm{graph}(G^{-1}) \setminus \{x = y\} \big) \cap P_1 \big( \textrm{graph}(G) \setminus \{x = y\} \big), \\
\label{E:TRe}
\mathcal T_e :=&~ P_1 \big( \textrm{graph}(G^{-1}) \setminus \{x = y\} \big) \cup P_1 \big( \textrm{graph}(G) \setminus \{x = y\} \big).
\end{align}
\end{subequations}
\end{definition}
From the definition of $G$ one can prove that $\mathcal{T}$, $\mathcal{T}_e$ are analytic sets.
The subscript $e$ refers to the endpoints of the geodesics: we have
\begin{equation}
\label{E:RTedef}
\mathcal{T}_e = P_1(R \setminus \{x = y\}).
\end{equation}
It follows that we have only to study the Monge problem in $\mathcal{T}_e$: $\pi(\mathcal{T}_e \times \mathcal{T}_e \cup \{x = y\}) = 1$.
As a consequence, $\mu(\mathcal{T}_e) = \nu(\mathcal{T}_e)$ and any
maps $T$ such that for $\nu \llcorner_{\mathcal{T}_e} = T_\sharp \mu \llcorner_{\mathcal{T}_e}$ can be extended to a
map $T'$ such that $\nu = T_\sharp \mu$ with the same cost by setting
\begin{equation}
\label{E:extere}
T'(x) =
\begin{cases}
T(x) & x \in \mathcal{T}_e \crcr
x & x \notin \mathcal{T}_e
\end{cases}
\end{equation}
By the non-branching assumption, if $x \in \mathcal{T}$, then $R(x)$ is a single geodesic and therefore
the set $R \cap \mathcal{T} \times \mathcal{T}$ is an equivalence relation on $\mathcal{T}$ that we will call ray equivalence relation.
Notice that the set $G$ is a partial order relation on $\mathcal{T}_e$.
The next step is to study the set $\mathcal{T}_{e}\setminus \mathcal{T}$.
\begin{definition}
\label{D:endpoint}
Define the multivalued \emph{endpoint graphs} by:
\begin{subequations}
\label{E:endpoint0}
\begin{align}
\label{E:endpointa}
a :=&~ \big\{ (x,y) \in G^{-1}: G^{-1}(y) \setminus \{y\} = \emptyset \big\}, \\
\label{E:endpointb}
b :=&~ \big\{ (x,y) \in G: G(y) \setminus \{y\} = \emptyset \big\}.
\end{align}
\end{subequations}
We call $P_2(a)$ the set of \emph{initial points} and $P_2(b)$ the set of \emph{final points}.
\end{definition}
Even if $a$, $b$ are not in the analytic class, still they belong to the $\sigma$-algebra $\mathcal{A}$.
\begin{proposition}
The following holds:
\begin{enumerate}
\item the sets
\[
a,b \subset X \times X, \quad a(A), b(A) \subset X,
\]
belong to the $\mathcal{A}$-class if $A$ analytic;
\item $a \cap b \cap \mathcal{T}_e \times X = \emptyset$;
\item $a(x)$, $b(x)$ are singleton or empty when $x \in \mathcal{T}$;
\item $a(\mathcal{T}) = a(\mathcal{T}_e)$, $b(\mathcal{T}) = b(\mathcal{T}_e)$;
\item $\mathcal{T}_e = \mathcal{T} \cup a(\mathcal{T}) \cup b(\mathcal{T})$, $\mathcal{T} \cap (a(\mathcal{T}) \cup b(\mathcal{T})) = \emptyset$.
\end{enumerate}
\end{proposition}
Finally we can assume that the $\mu$-measure of final points and the $\nu$-measure of the initial points are $0$:
indeed since the sets $G \cap b(\mathcal{T}) \times X$, $G \cap X \times a(\mathcal{T})$ is a subset of the graph of the identity map,
it follows that from the definition of $b$ one has that
\[
x \in b(\mathcal{T}) \quad \Longrightarrow \quad G(x) \setminus \{x\} = \emptyset,
\]
A similar computation holds for $a$. Hence we conclude that
\[
\pi (b(\mathcal{T}) \times X) = \pi(G \cap b(\mathcal{T}) \times X) = \pi(\{x = y\}),
\]
and following \eqref{E:extere} we can assume that
\[
\mu(b(\mathcal{T})) = \nu(a(\mathcal{T})) = 0.
\]
\subsection{The Wiener case}
\label{Ss:Wiener}
For the abstract Wiener space, it is possible to obtain more regularity for the sets introduced so far.
Let $d= \| \cdot \|$ and $d_{L} = \| \cdot \|_{H}$: by the compactness of the embedding $R_{\gamma}$ of $H$ into X it follows that
\begin{itemize}
\item[(1)] $d_L : X \times X \to [0,+\infty]$ l.s.c. distance;
\item[(2)] $d_L(x,y) \geq C d(x,y)$ for some positive constant $C$;
\item[(3)] $\cup_{x \in K_1, y \in K_2} \gamma_{[x,y]}$ is $d$-compact if $K_1$, $K_2$ are $d$-compact, $d_L \llcorner_{K_1 \times K_2}$ uniformly bounded.
\end{itemize}
The set $\Gamma'$ is $\sigma$-compact: in fact, if one restrict to each $\Gamma_n$
given by \eqref{E:gamman}, then the set of cycles of order $I$ is compact, and thus
\begin{align*}
\Gamma'_{n,\bar I} :=&~ \bigg\{ (x,y) : \exists I \in \{0,\dots,\bar I\}, (w_i,z_i) \in \Gamma_n \ \text{for} \ i = 0,\dots,I, \ z_I = y \crcr
&~ \qquad \qquad w_{I+1} = w_0 = x, \ \sum_{i=0}^I d_L(w_{i+1},z_i) - d_L(w_i,z_i) = 0 \bigg\}
\end{align*}
is compact. Finally $\Gamma' = \cup_{n,I} \Gamma'_{n,I}$.
Moreover, $d_L \llcorner_{\Gamma'_{n,I}}$ is continuous. If $(x_n,y_n) \to (x,y)$, then from the l.s.c. and
\[
\sum_{i=0}^I d_L(w_{n,i+1},z_{n,i}) = \sum_{i=0}^I d_L(w_{n,i},z_{n,i}), \quad w_{n,I+1} = w_{n,0} = x_{n}, \ z_{n,I} = y_{n},
\]
it follows also that each $d_L(w_{n,i+1},z_{n,i})$ is continuous.
Similarly the sets $G$, $R$, $a$, $b$ are $\sigma$-compact: assumption (3) and the above computation in fact shows that
\[
G_{n,I} := \Big\{ (x,y): \exists (w,z) \in \Gamma'_{n,I}, d_L(w,x) + d_L(x,y) + d_L(y,z) = d_L(w,z) \Big\}
\]
is compact. For $a$, $b$, one uses the fact that projection of $\sigma$-compact sets is $\sigma$-compact.
So we have that $\Gamma$, $\Gamma'$, $G$, $G^{-1}$, $a$ and $b$ are $\sigma$-compact sets.
\subsection{Strongly consistency of disintegrations}\label{Ss:partition}
We recall the main results of \cite{biacava:streconv} that permit to define an order preserving map $g$ which maps our transport problem into a transport problem on $\mathcal S \times \mathbb{R}$, where $\mathcal S$ is a cross section of $R$.
The strong consistency of the disintegration follows from the next result.
\begin{proposition}
\label{P:sicogrF}
There exists a $\mu$-measurable cross section $f : \mathcal{T} \to \mathcal{T}$ for the ray equivalence relation $R$.
\end{proposition}
Up to a $\mu$-negligible saturated set $\mathcal{T}_N$, we can assume it to have $\sigma$-compact range: just let $S \subset f(\mathcal{T})$ be a $\sigma$-compact set where
$f_\sharp \mu \llcorner_{\mathcal{T}}$ is concentrated, and set
\begin{equation}
\label{E:TNngel}
\mathcal{T}_S := R^{-1}(S) \cap \mathcal{T}, \quad \mathcal{T}_N := \mathcal{T} \setminus \mathcal{T}_S, \quad \mu(\mathcal{T}_N) = 0.
\end{equation}
Having the $\mu\llcorner_{\mathcal{T}}$-measurable cross-section
\[
\mathcal{S} := f(\mathcal{T})= S \cup f(\mathcal{T}_N) = (\textrm{Borel}) \cup (f(\text{$\mu$-negligible})),
\]
we can define the parametrization of $\mathcal{T}$ and $\mathcal{T}_e$ by geodesics.
Using the quotient map $f$, we obtain a unitary speed parametrization of the transport set.
\begin{definition}[Ray map]
\label{D:mongemap}
Define the \emph{ray map $g$} by the formula
\begin{align*}
g :=&~ \Big\{ (y,t,x): y \in \mathcal{S}, t \in [0,+\infty), x \in G(y) \cap \{d_L(x,y) = t\} \Big\} \crcr
&~ \cup \Big\{ (y,t,x): y \in \mathcal{S}, t \in (-\infty,0), x \in G^{-1}(y) \cap \{d_L(x,y) = -t\} \Big\} \crcr
=&~ g^+ \cup g^-.
\end{align*}
\end{definition}
\begin{proposition}
\label{P:gammaclass}
The following holds.
\begin{enumerate}
\item The restriction $g \cap S \times \mathbb{R} \times X$ is analytic.
\item The set $g$ is the graph of a map with range $\mathcal{T}_e$.
\item $t \mapsto g(y,t)$ is a $d_L$ $1$-Lipschitz $G$-order preserving for $y \in \mathcal{T}$.
\item $(t,y) \mapsto g(y,t)$ is bijective on $\mathcal{T}$, and its inverse is
\[
x \mapsto g^{-1}(x) = \big( f(y),\pm d_L(x,f(y)) \big)
\]
where $f$ is the quotient map of Proposition \ref{P:sicogrF} and the positive/negative sign depends on $x \in G(f(y))$/$x \in G^{-1}(f(y))$.
\end{enumerate}
\end{proposition}
Another property of $d_{L}$-cyclically monotone transference plans.
\begin{proposition}
\label{P:ortho}
For any $\pi$ $d_{L}$-monotone there exists a $d_L$-cyclically monotone transference plan $\tilde \pi$ with the same cost of $\pi$ such that it coincides with the identity on $\mu \wedge \nu$.
\end{proposition}
Coming back to the abstract Wiener space, we have that given $\mu,\nu \ll\gamma$ and given $\pi \in \Pi(\mu,\nu)$ $\| \cdot \|_{H(\gamma)}$-cyclically monotone,
we have constructed the transport $\mathcal{T}$ (and $\mathcal{T}_{e}$), an equivalence relation $R$ on it with geodesics as equivalence classes
and the corresponding disintegration
is strongly consistent:
\begin{equation}\label{E:primadis}
\mu\llcorner_{\mathcal{T}} = \int_{\mathcal{S}} \mu_{y} m(dy)
\end{equation}
with $m = f_{\sharp} \mu$ and $\mu_{y}(R(y)) = 1$ for $m$-a.e. $y \in \mathcal{T}$. Using the ray map $g$ one can assume that $\mu_{y} \in \mathcal{P}(\mathbb{R})$ and
\[
\mu\llcorner_{\mathcal{T}} = g_{\sharp} \int_{\mathcal{S}} \mu_{y} m(dy) .
\]
\section{Regularity of disintegration}
\label{S:disin}
To obtain existence of an optimal transport map it is enough to prove that $\mu$ in concentrated on $\mathcal{T}$ and
$\mu_{y}$ is a continuous measure for $m$-a.e. $y \in \mathcal{S}$. Indeed at that point, for every $y \in \mathcal{S}$ we consider
the unique monotone map $T_{y}$ such that $T_{y\,\sharp}\mu_{y}=\nu_{y}$, then $T(g(y,t)): = T_{y}(g(y,t))$ is an optimal transport map,
see Theorem 6.2 of \cite{biacava:streconv}.
In Section \ref{S:disin} we introduce the fundamental regularity assumption (Assumption \ref{A:NDEatom}) on the measure $\gamma$
and we show that it implies the $\gamma$-negligibility of the set of initial points. Consequently we obtain a disintegration of $\mu$
on the whole space.
From Assumption \ref{A:NDEatom} it also follows that the disintegration of $\mu$ w.r.t. the ray equivalence relation $R$
has continuous conditions probabilities.
Define the map $ X\times X \ni (x,y) \mapsto T_{t}(x,y) : = x (1-t ) + y t$.
\begin{assumption}[Non-degeneracy assumption]
\label{A:NDEatom} The measure $\gamma$ is said to satisfy Assumption \ref{A:NDEatom} w.r.t. a $\| \cdot \|_{H(\gamma)}$-cyclically monotone set
$\Gamma$ if
\begin{itemize}
\item[i)] $\pi(\Gamma)=1$ with $\pi \in \Pi(\mu,\nu)$ and $\mu,\nu \ll \gamma$;
\item [ii)] for each closed set $A$ such that $\mu (A) > 0$ there exists $C >0 $ and $\{ t_{n} \}_{n\in \mathbb{N}} \subset [0,1]$ converging to $0$ as
$n\to +\infty$ such that
\[
\gamma( T_{t_{n}}(\Gamma \cap A\times X) ) \geq C \mu (A)
\]
for all $n \in \mathbb{N}$.
\end{itemize}
\end{assumption}
Clearly it is enough to verify Assumption \ref{A:NDEatom} for $A$ compact set.
An immediate consequence of the Assumption \ref{A:NDEatom} is that the final points are $\gamma$-negligible.
\begin{proposition}
\label{P:puntini}
If $\gamma$ satisfies Assumption \eqref{A:NDEatom} then
\[
\mu( a(\mathcal{T}_e)) = 0.
\]
\end{proposition}
\begin{proof}
Let $A= a(\mathcal{T}_{e})$. Suppose by contradiction $\mu( A)>0$.
By inner regularity and $\Gamma \subset \{ (x,y) : \|x-y\|_{H(\gamma)}< + \infty \}$,
there exists a compact set $\hat A \subset A$ such that $\mu (\hat A)>0$ and $\rho_{1}(x)\geq \delta$ for all $x\in \hat A$ and for some constant $\delta>0$.
Moreover we can assume that
\[
\Gamma \cap \hat A \times X \subset \{ (x,y) : \|x-y\|_{H(\gamma)}\leq M \}
\]
for some positive $M \in \mathbb{R}$.
By Assumption \ref{A:NDEatom} there exist $C > 0$ and
$\{ t_{n} \}_{n\in \mathbb{N}}$ converging to 0 such that
\[
\gamma( T_{t_{n}}(\Gamma \cap \hat A \times X) ) \geq C \mu ( \hat A ) \geq \delta C \gamma(\hat A).
\]
Denote with $\hat A_{t_{n}} = T_{t_{n}}(\Gamma \cap \hat A \times X)$ and define $ \hat A^{\varepsilon}:= \big\{ x : \| \hat A -x\|_{H(\gamma)} < \varepsilon \big\}$.
Since $\hat A \subset A = a(\mathcal{T}_{e})$, $\hat A_{t_{n}} \cap \hat A = \emptyset$ for every $n \in \mathbb{N}$.
Moreover for $t_{n}\leq\varepsilon/M$ it holds $\hat A^{\varepsilon} \supset \hat A_{t_{n}}$. So we have for $t_{n}$ small enough
\[
\gamma(\hat A^{\varepsilon}) \geq \gamma(\hat A) + \gamma(\hat A_{t_{n}}) \geq (1+ C\delta) \gamma (\hat A).
\]
Since $\gamma(\hat A) = \lim_{\varepsilon \to 0} \gamma (\hat A^{\varepsilon})$, this is a contradiction.
\end{proof}
It follows that $\mu(\mathcal{T}) = 1$, therefore we can use the Disintegration Theorem \ref{T:disintr} to write
\begin{equation}
\label{E:disintT}
\mu = \int_S \mu_y m(dy), \quad m = f_\sharp \mu, \ \mu_y \in \mathcal{P}(R(y)).
\end{equation}
The disintegration is strongly consistent since the quotient map $f : \mathcal T \to \mathcal T$ is $\mu$-measurable and $(\mathcal T,\mathcal B(\mathcal T))$ is countably generated.
The second consequence of Assumption \ref{A:NDEatom} is that $\mu_y$ is continuous, i.e. $\mu_y(\{x\}) = 0$ for all $x \in X$.
\begin{proposition}
\label{P:nonatoms}
If $\gamma$ satisfies Assumption \ref{A:NDEatom} then the conditional probabilities $\mu_y$ are continuous for $m_{\gamma}$-a.e. $y \in S$.
\end{proposition}
\begin{proof}
From the regularity of the disintegration and the fact that $m(S) = 1$, we can assume that the map $y \mapsto \mu_y$ is weakly continuous on a
compact set $K \subset S$ of comeasure $<\varepsilon$.
It is enough to prove the proposition on $K$.
{\it Step 1.} From the continuity of $K \ni y \mapsto \mu_y \in \mathcal{P}(X)$ w.r.t. the weak topology, it follows that the map
\[
y \mapsto A(y) := \big\{ x \in R(y): \mu_y(\{x\}) > 0 \big\} = \cup_n \big\{ x \in R(y): \mu_y(\{x\}) \geq 2^{-n} \big\}
\]
is $\sigma$-closed: in fact, if $(y_m,x_m) \to (y,x)$ and $\mu_{y_m}(\{x_m\}) \geq 2^{-n}$, then $\mu_y(\{x\}) \geq 2^{-n}$ by u.s.c. on compact sets.
Hence $A$ is Borel.
{\it Step 2.} The claim is equivalent to $\mu( P_{2}(A))=0$. Suppose by contradiction $\mu(P_{2}(A))>0$.
By Lusin Theorem (Theorem 5.8.11 of \cite{Sri:courseborel})
$A$ is the countable union of Borel graphs.
Therefore we can take a Borel selection of $A$ just considering one of the Borel graphs, say $\hat A$.
Clearly $m (P_{1}(\hat A)) > 0$ hence by \eqref{E:disintT} $\mu(P_{2}(\hat A))>0$.
By Assumption \ref{A:NDEatom} $\gamma( T_{t_{n}}(\Gamma \cap P_{2}(\hat A) \times X) ) \geq C \mu(P_{2}(\hat A))$ for some $C>0$ and $t_{n}\to 0$.
From $T_{t_{n}}(\Gamma \cap P_{2}(\hat A) \times X) \cap (P_{2}(\hat A)) = \emptyset$, using the same argument of Proposition \ref{P:puntini}, the claim follows.
\end{proof}
\section{An approximation result}
\label{S:approx}
In Section \ref{S:disin} we proved that if $\gamma$ satisfies Assumption \ref{A:NDEatom} w.r.t. a $\| \cdot \|_{H(\gamma)}$-cyclically monotone set
$\Gamma$ such that $\pi(\Gamma)=1$, then the disintegration of $\mu$ has enough regularity to solve the Monge minimization problem.
The remaining part of the paper will be devoted to proving that the centred non-degenerate Gaussian measure $\gamma$
verifies Assumption \ref{A:NDEatom} w.r.t. a $\| \cdot \|_{H(\gamma)}$-cyclically monotone set
$\Gamma$ such that $\pi(\Gamma)=1$ with $\pi \in \Pi(\mu,\nu)$ and $\mu,\nu \ll \gamma$.
In particular in this section we prove that it is enough to verify Assumption \ref{A:NDEatom} for a well prepared finite dimensional approximation,
provided some uniformity holds.
Let $P_{d} : X \to H$ be the projection map of Proposition \ref{P:appro} associated to the orthonormal basis $\{e_{i}\}_{i \in \mathbb{N}}$ of $H(\gamma)$
with $e_{i} = R_{\gamma} \hat e_{i}$ for $\hat e_{i} \in R_{\gamma}^{*}X^{*}$ and $P_{d\,\sharp} \gamma = \gamma_{d}$.
Consider the following measures
\begin{equation}\label{E:approxvera}
\mu_{d} : = P_{d\,\sharp} \mu, \qquad \nu_{d} : = P_{d\,\sharp} \nu
\end{equation}
and observe that $\mu_{d}= \rho_{1,d} \gamma_{d}$ and $\nu_{d}= \rho_{2,d} \gamma_{d}$
with
\begin{equation}\label{E:densita}
\rho_{i,d}(z) = \int \rho_{i}(x) \gamma_{z,d}^{\perp} (dx), \qquad i =1,2,
\end{equation}
where $\gamma_{z,d}$ is defined in \ref{E:disintg}. Recall that $\mu_{d} \rightharpoonup \mu$ and $\nu_{d} \rightharpoonup \nu$ as $d \nearrow + \infty$.
Since $\rho_{i,d}$ depend only on the first $d$-coordinates, the measures $\mu_{d},\nu_{d}$
can be considered as probability measure on $\mathbb{R}^{d}$. Therefore
we can study the transport problem with euclidean norm cost $\|x \|_{d}^{2} : = \sum_{j=1}^{d} x_{j}^{2}$:
\begin{equation}\label{E:traspapprox}
\min_{\pi \in \Pi(\mu_{d},\nu_{d})} \int \|x-y \|_{d} \pi(dx dy).
\end{equation}
It is a well-known fact in optimal transportation that this problem has a minimizer of the form $(Id,T_{d})_{\sharp}\mu_{d}$
with $T_{d}$ invertible and Borel.
For each $d$ we choose as optimal map $T_{d}$ the one obtained gluing the monotone rearrangements over the geodesics.
\begin{proposition}\label{P:limiteottimo}
Let $\pi_{d} \in \Pi(\mu_{d},\nu_{d})$ be an optimal transference plan for \eqref{E:traspapprox}
such that $\pi_{d} \rightharpoonup \pi$.
Then $\pi \in \Pi(\mu,\nu)$ is an optimal transport plan for \eqref{E:trasp}.
\end{proposition}
\begin{proof}
Let $\hat \pi \in \Pi(\mu,\nu)$ be an optimal transference plan. The following holds true
\begin{align*}
\int \|x-y \|_{H} \hat \pi(dxdy) \geq &~ \int \|P_{d}(x-y) \|_{H(\gamma)} \hat \pi(dxdy) = \int \|x-y \|_{H(\gamma)} ((P_{d}\otimes P_{d})_{\sharp}\hat \pi)(dxdy) \crcr
\geq &~ \int \|x-y \|_{H(\gamma)} \pi_{d}(dxdy).
\end{align*}
Since $\| \cdot \|_{H(\gamma)}$ is l.s.c. and $\pi_{d} \rightharpoonup \pi$ it follows that
\[
\int \|x-y \|_{H(\gamma)} \hat \pi(dxdy) \geq \liminf_{d \to + \infty} \int \|x-y \|_{H(\gamma)} \pi_{d}(dxdy) \geq \int \|x-y \|_{H(\gamma)}\pi(dxdy).
\]
Hence the claim follows.
\end{proof}
If the sequence $\pi_{d} \in \Pi(\mu_{d},\nu_{d})$ of optimal transference plans
satisfies, for every $d \in \mathbb{N}$, Assumption \ref{A:NDEatom} w.r.t. $\Gamma_{d} = \textrm{graph}(T_{d})$ with $C$ independent on $d$,
then the optimal transference plan $\pi$, weak limit of $\pi_{d}$, satisfies Assumption \ref{A:NDEatom} w.r.t. a $\| \cdot \|_{H(\gamma)}$-cyclically monotone set $\Gamma$ such that $\pi(\Gamma)=1$.
\begin{theorem}\label{T:approx}
Assume that there exists $C > 0$ such that for all $d \in \mathbb{N}$ and for all $A \subset X$ compact set
the following holds true
\[
\gamma_{d} \big( T_{t}( \Gamma_{d} \cap A \times X) \big) \geq C \mu_{d}( A).
\]
Then for all $A \subset X$ compact set
\begin{equation}\label{E:stima}
\gamma \big( T_{t}( \Gamma \cap A \times X)\big) \geq C \mu( A ),
\end{equation}
where $\Gamma \subset X \times X$ is $\| \cdot \|_{H(\gamma)}$-cyclically monotone with $\pi(\Gamma)=1$.
\end{theorem}
\begin{proof}
Assume that $A = C(B) \in \mathcal{E}(X)$ with $B \in \mathbb{R}^{n}$ compact set for some fixed $n \in \mathbb{N}$.
It follows from Proposition \ref{P:limiteottimo}
that $\pi$ is an optimal transference plan concentrated on a $\| \cdot \|_{H(\gamma)}$-cyclically monotone set $\Gamma$.
Observe that $\Gamma_{d}$ is closed for every $d \in \mathbb{N}$ .
{\it Step 1.}
Since $\mu_{d} \rightharpoonup \mu$ and $\nu_{d} \rightharpoonup \nu$, for every $\varepsilon >0$ there exist $K_{1,\varepsilon}$ and $K_{2,\varepsilon}$ compact sets such that
$\mu_{d}(K_{1,\varepsilon}) \geq 1- \varepsilon/2$ and $\nu_{d}(K_{2,\varepsilon}) \geq 1- \varepsilon/2$.
Consider the compact se $\Gamma_{d,\varepsilon} := \Gamma_{d} \cap K_{1,\varepsilon} \times K_{2,\varepsilon}$,
then $\pi_{d}(\Gamma_{d,\varepsilon}) \geq 1-\varepsilon$ and $\Gamma_{d,\varepsilon}$ converges in the Hausdorff topology,
up to subsequences for $d \to + \infty$, to a set $\Gamma_{\varepsilon} \subset \Gamma$ with $\pi(\Gamma_{\varepsilon})\geq 1-\varepsilon$.
Since $T_{t}$ is continuous, $T_{t}(\Gamma_{d,\varepsilon} \cap A \times X)$ is compact and by
\[
T_{t}(\Gamma_{d,\varepsilon} \cap A\times X) \subset \overline{co} (P_{1}(K_{\varepsilon} \cap A\times X) \cap P_{2}(K_{\varepsilon} \cap A\times X))
\]
it follows that $T_{t}(\Gamma_{d,\varepsilon} \cap A\times X)$ converges in the Hausdorff topology to $T_{t}(\Gamma_{\varepsilon} \cap A\times X)$.
{\it Step 2.}
It follows that
\[
\gamma(T_{t}(\Gamma_{\varepsilon} \cap A \times X)) \geq \limsup_{d \to + \infty} \gamma_{d}(T_{t}(\Gamma_{d,\varepsilon} \cap A \times X)),
\]
hence, using the fact that $\Gamma_{d,\varepsilon}$ is a subset of a graph, it follows that
\begin{align}\label{E:bene}
\gamma(T_{t}(\Gamma_{\varepsilon} \cap A \times X)) \geq &~ \limsup_{d \to + \infty} \gamma_{d}(T_{t}(\Gamma_{d,\varepsilon} \cap A \times X))\crcr
\geq &~ C \limsup_{d \to + \infty} \mu_{d} ( P_{1}(\Gamma_{d,\varepsilon}) \cap A) \crcr
\geq &~ C \limsup_{d \to + \infty} \mu_{d} (A) - C \varepsilon
\end{align}
where in the last equation we have used $\mu_{d} (P_{1}(\Gamma_{d,\varepsilon})) \geq 1-\varepsilon$.
{\it Step 3.} Since $\mu_{d}= P_{d\,\sharp} \mu$
\begin{equation}\label{E:bbene}
\gamma(T_{t}(\Gamma_{\varepsilon} \cap A \times X)) \geq C \limsup_{d \to + \infty} \mu_{d} (A) - C \varepsilon =
C\mu(A) - C\varepsilon.
\end{equation}
for all $A \in \mathcal{E}(X)$ with finite dimensional base; note that this family of sets is a base for the weak topology.
It follows from the compactness of $\Gamma_{\varepsilon}$ that \eqref{E:bbene} is stable under uncountable intersection,
hence \eqref{E:bbene} holds true for all weak closed subset of $\Gamma_{\varepsilon}$. Since $\Gamma_{\varepsilon}$ is a compact set,
weak topology has the same closed set of the strong one. It follows that for all closed set $A \subset X$
\[
\gamma(T_{t}(\Gamma \cap A \times X)) \geq \gamma(T_{t}(\Gamma_{\varepsilon} \cap A \times X)) \geq C \mu ( A) - C \varepsilon.
\]
Passing to the limit as $\varepsilon \to 0$, the claim follows.
\end{proof}
\section{Finite dimensional estimate}
\label{S:finite}
Throughout this Section we will use the following notation $Jac(F)(x) : = |\det dF|(x)$ for any map $F:\mathbb{R}^{d} \to \mathbb{R}^{d}$.
In the next theorem we prove that $d$-dimensional standard Gaussian measure $\gamma_{d} = P_{d\,\sharp} \gamma$ satisfies
Assumption \ref{A:NDEatom} for $\Gamma = \textrm{graph}(T_{d})=\Gamma_{d}$:
\[
\gamma_{d} \big( T_{t}( \Gamma_{d} \cap A \times X) \big) \geq C \mu_{d}( A ).
\]
Observe that the set $T_{t}( \Gamma_{d} \cap A \times X)$ is parametrized by the map $T_{d,t}:= Id(1-t) + T_{d} t$.
\begin{theorem}\label{T:stimapprox}
Assume that there exists $M>0$ such that
$\rho_{i,d} (x) \leq C$ for $\gamma_{d}$-a.e. $x \in \mathbb{R}^{d}$ and $i=1,2$.
Then the following estimate holds true
\[
\gamma_{d} \big( T_{d,t}( A ) \big) \geq \frac{1}{C} \mu_{d}(A ).
\]
\end{theorem}
\begin{proof} During the proof we will omit the subscript $d$.
{\it Step 1.} Consider the Monge minimization problem with cost $c_{p}$, \eqref{E:monge},
between $\mu_{d}$ and $\nu_{d}$. It follows from Theorem \ref{T:different} and the boundedness of $\rho_{i,d}$ that
there exists a unique optimal map $T_{p}$ approximately differentiable $\mu_{d}$-a.e., hence
by Lemma \ref{L:density} it follows that
\[
\rho_{2}(T_{p}(x)) |\det \tilde \nabla T_{p}|(x)\prod_{j=1}^{d}\frac{1}{\sqrt{2\pi}} \exp\Big\{ - \frac{T_{p}(x)_{j}^{2}}{2} \Big\} =
\rho_{1}(x) \prod_{j=1}^{d}\frac{1}{\sqrt{2\pi}} \exp\Big\{ - \frac{x_{j}^{2}}{2} \Big\}.
\]
Since for $\mu_{d}$-a.e. $x \in\mathbb{R}^{d}$ $|\det \tilde \nabla T_{p}|(x)>0$, also $\rho_{2}(T_{p}(x))>0$ for $\mu_{d}$-a.e. $x \in \mathbb{R}^{d}$.
Hence the following makes sense $\mu$-a.e.:
\[
Jac(T_{p})(x)= |\det \tilde \nabla T_{p}|(x) = \frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))} \exp \Big\{ \sum_{j=1}^{d} - \frac{1}{2} ( x_{j}^{2} - T_{p}(x)_{j}^{2} ) \Big\}.
\]
{\it Step 2.} Let $T_{p,t} : = Id (1-t) + T_{p} t$.
From Theorem \ref{T:different}, $\det \tilde \nabla T_{p}(x) = \prod_{j=1}^{d}\lambda_{j}$ with $\lambda_{i}>0$ for $i=1,\dots,d$.
It follows that
\[
Jac(T_{p,t})(x) = \det (Id (1-t) + \tilde\nabla T_{p}(x)t) = \prod_{j=1}^{d}\big( (1-t ) + \lambda_{j}t \big).
\]
Passing to logarithms, we have by concavity
\[
\log (Jac(T_{p,t})(x) ) \geq t \log (Jac(T_{p})(x) ) \Longrightarrow Jac(T_{p,t})(x) \geq (Jac(T_{p})(x))^{t}.
\]
Hence
\begin{equation}
Jac(T_{p,t})(x) \geq \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t} \exp \Big\{ \sum_{j=1}^{d} - \frac{1}{2} t ( x_{j}^{2} - T_{p}(x)_{j}^{2} ) \Big\}.
\end{equation}
{\it Step 3.}
We have the following
\begin{align*}
\exp\Big\{ \sum_{j=1}^{d}- &~\frac{1}{2} (T_{p,t}(x)_{j}^{2} - x_{j}^{2}) \Big\} Jac (T_{p,t})(x) \crcr
&~ \geq \exp\Big\{ \sum_{j=1}^{d}- \frac{1}{2} (T_{p,t}(x)_{j}^{2} - x_{j}^{2}) \Big\}
\Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t} \exp \Big\{ \sum_{j=1}^{d} - \frac{1}{2} t ( x_{j}^{2} - T_{p}(x)_{j}^{2} ) \Big\} \crcr
&~ = \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t}
\exp\Big\{ \sum_{j=1}^{d}- \frac{1}{2} (T_{p,t}(x)_{j}^{2} - x_{j}^{2} +t x_{j}^{2} - t T_{p}(x)_{j}^{2} ) \Big\} \crcr
&~ = \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t}
\exp\Big\{ \sum_{j=1}^{d}- \frac{1}{2} \Big( ((1-t)x_{j} + t T_{p}(x)_{j})^{2} - ( (1-t) x_{j}^{2} + t T_{p}(x)_{j}^{2} \Big) \Big\} \crcr
&~ = \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t}
\exp\Big\{ \sum_{j=1}^{d}- \frac{1}{2} \big( x_{j} -T_{p}(x)_{j} \big)^{2} (t^{2}-t) \Big\} \crcr
&~ = \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t}
\exp\Big\{ - \frac{1}{2} \| x -T_{p}(x) \|_{d}^{2} (t^{2}-t) \Big\}.
\end{align*}
Hence
\begin{align*}
\gamma(T_{p,t}(A)) = &~ \int_{A} Jac(T_{p,t})(x) \prod_{j=1}^{d}\frac{1}{\sqrt{2\pi}} \exp\Big\{ - \frac{1}{2} T_{p,t}(x)_{j}^{2} \Big\} \mathcal{L}^{d}(dx) \crcr
= &~ \int_{A} Jac(T_{p,t})(x) \exp\Big\{ \sum_{j=1}^{d} - \frac{1}{2} ( T_{p,t}(x)_{j}^{2} - x_{j}^{2} ) \Big\} \gamma(dx) \crcr
\geq &~ \int_{A} \Big(\frac{\rho_{1}(x)}{\rho_{2}(T_{p}(x))}\Big)^{t} \exp\Big\{ \frac{1}{2} \| x -T_{p}(x) \|_{d}^{2} (t- t^{2}) \Big\} \gamma(dx) \crcr
\geq &~ \frac{1}{C^{t}}\int_{A} \rho_{1}(x)^{t} \gamma(dx) \crcr
\geq &~ \frac{1}{C^{t}}\int_{A}\rho_{1}(x)^{t-1} \mu(dx) \crcr
\geq &~ \frac{1}{C}\mu(A).
\end{align*}
The claim follows.
{\it Step 4.} Since $(Id,T_{p})_{\sharp}\mu_{d} \rightharpoonup (Id,T)_{\sharp} \mu_{d}$ as $p \searrow 1$, using the same techniques of Theorem \ref{T:approx}'s proof,
it is fairly easy to prove that
\[
\gamma_{d}(T_{t}(A)) \geq \frac{1}{C} \mu_{d}( A).
\]
\end{proof}
\begin{remark}\label{R:resume}
We summarize the results obtained so far. If $\rho_{1},\rho_{2} \leq M$,
then from \eqref{E:densita} it follows that the densities of $\mu_{d}$ and $\nu_{d}$ enjoy the same property with the same constant $M$.
Hence using the approximating sequences $\mu_{d}$ and
$\nu_{d}$, we have from Theorem \ref{T:approx} and Theorem \ref{T:stimapprox} that
\[
\gamma \big( T_{t}( \Gamma \cap A \times X)\big) \geq \frac{1}{M} \mu( A).
\]
As Proposition \ref{P:puntini} and Proposition \ref{P:nonatoms} show, this estimate implies $\mu(a(\mathcal{T}))=0$ and the continuity of the conditional
probabilities $\mu_{y}$.
Since the optimal finite dimensional map $T_{d}$ is invertible, following the argument of Theorem \ref{T:stimapprox} we can also prove
\begin{equation}\label{E:ricapitolo}
\gamma \big( T_{1-t}( \Gamma \cap X \times A)\big) \geq \frac{1}{M} \nu( A),
\end{equation}
and adapting the proofs of Proposition \ref{P:puntini} and Proposition \ref{P:nonatoms} we can prove that $\nu(b(\mathcal{T}))=0$ and
the continuity of the conditional probabilities $\nu_{y}$. So we have
\[
\mu = \int_{\mathcal{S}} \mu_{y} m (dy), \quad \nu = \int_{\mathcal{S}} \nu_{y} m (dy), \qquad \mu_{y}, \nu_{y} \ \textrm{continuous for } m-a.e. y \in \mathcal{S}.
\]
In the next Section we remove the hypothesis $\rho_{1},\rho_{2} \leq M$.
\end{remark}
\section{Solution}
\label{S:solu}
In this Section we obtain the existence of an optimal transport map for the Monge minimization problem between measures $\mu$ and $\nu$
absolute continuous w.r.t. $\gamma$. We first prove that the set of initial points and the set of final points have $\mu$-measure zero and $\nu$-measure zero respectively.
Recall that the initial points map $a$ and the final points map $b$ have been introduced in Definition \ref{D:endpoint}.
\begin{proposition}\label{P:noiniz}
Let $\mu, \nu \in \mathcal{P}(X)$ be such that $\mu,\nu \ll \gamma$. Then $\mu(a(\mathcal{T}))= \nu(b(\mathcal{T}))=0$.
\end{proposition}
\begin{proof} Let $\mu = \rho_{1}\gamma$ and $\nu =\rho_{2}\gamma$.
We prove that $\mu(a(\mathcal{T}))= 0$.
{\it Step 1.} Assume by contradiction that $\mu(a(\mathcal{T}))>0$.
Let $A \subset a(\mathcal{T})$ be such that $\mu(A)>0$ and for every $x \in A$,
$\rho_{1}(x)\leq M$ for some positive constant $M$.
Consider $\gamma\llcorner_{\mathcal{T}}$ and its disintegration
\[
\gamma\llcorner_{\mathcal{T}} = \int_{\mathcal{S}} \gamma_{y} m_{\gamma}(dy), \quad \gamma_{y}(\mathcal{T})=1, \ m_{\gamma}-a.e. y \in \mathcal{S}.
\]
Consider the initial point map $a : \mathcal{S} \to A$ and the measure $a_{\sharp} m_{\gamma}$. Observe that since
\[
\forall B \subset A: \mu(B)>0 \quad \Rightarrow \quad \gamma( R(B) \cap \mathcal{T})>0,
\]
it follows that $\mu\llcorner_{A} \ll a_{\sharp} m_{\gamma}$. Hence there exists $\hat A \subset A$ of positive $a_{\sharp}m_{\gamma}$-measure such that
the map
\[
\hat A\ni x \mapsto h(x) : = \frac{d \mu\llcorner_{A}}{d a_{\sharp}m_{\gamma}} (x)
\]
verifies $ h(x) \leq M'$ for some positive constan $M'$.
{\it Step 2.}
Considering
\[
\mu\llcorner_{\hat A}, \qquad \hat \gamma : = \int_{R(\hat A) \cap \mathcal{S}} h(a(y))\gamma_{y} m_{\gamma}(dy),
\]
we have the claim. Indeed both have uniformly bounded densities w.r.t. $\gamma$ and $\mathcal{T}_{e}$ is still a transport set
for the transport problem between $\mu\llcorner_{\hat A}$ and $\hat \gamma$.
Indeed for $S \subset \mathcal{S}$
\begin{align*}
\mu\llcorner_{\hat A}(\cup_{y\in S} R(y)) = &~ \mu\llcorner_{\hat A} (a(S)) \crcr
= &~ \int_{a(S)} h(a) (a_{\sharp}m_{\gamma})(da) \crcr
= &~ \int_{S} h(a(y)) m_{\gamma}(dy) = \hat \gamma (\cup_{y\in S} R(y)).
\end{align*}
Hence we can project the measures, obtain the finite dimensional
estimate of Theorem \ref{T:stimapprox}, obtain the infinite dimensional estimate through Theorem \ref{T:approx} and finally by Proposition
\ref{P:puntini} get that $\mu(\hat A)=0$, that is a contradiction with $\mu(\hat A)>0$. In the same way, following Remark \ref{R:resume},
we obtain that $\nu(b(\mathcal{T}))=0$.
\end{proof}
It follows that the disintegration formula \eqref{E:primadis}
holds true on the whole transportation set:
\[
\mu = \int \mu_{y} m(dy), \qquad \nu = \int \nu_{y} m(dy).
\]
\begin{proposition}\label{P:noatom}
For $m$-a.e. $y \in \mathcal{S}$ the conditional probabilities $\mu_{y}$ and $\nu_{y}$ have no atoms.
\end{proposition}
\begin{proof} We only prove the claim for $\mu_{y}$.
{\it Step 1.} Suppose by contradiction that there exist a measurable set $\hat{\mathcal{S}} \subset \mathcal{S}$
such that $m(\hat{ \mathcal{S}})>0$
and for every $y \in \hat{ \mathcal{S}}$ there exists $x(y)$ such that $\mu_{y}(\{x(y)\})>0$.
Restrict and normalize both $\mu$ and $\nu$ to $R(\hat{\mathcal{S}})$, and denote them again with $\mu$ and $\nu$.
Consider the sets $K_{i,M} : = \{ x \in X : \rho_{i}\leq M \}$ for $i = 1,2$.
Note that
$\mu(K_{1,M}) \geq 1 -c_{1}(M)$ and $\nu(K_{2,\delta}) \geq 1 -c_{2}(M)$ with $c_{i}(M) \to 0$ as $M \nearrow +\infty$.
Hence for $M$ sufficiently large the conditional probabilities of the disintegration of $\mu\llcorner_{K_{1,M}}$
have atoms, therefore we can assume, possibly restricting $\hat{\mathcal{ S}}$, that for all $y \in \hat{\mathcal{S}}$ it holds $x(y) \in K_{1,M}$.
{\it Step 2.}
Define
\[
\mu_{y,M} : = \mu_{y}\llcorner_{K_{1,M}}, \qquad \nu_{y,M} : = \nu_{y}\llcorner_{K_{2,M}},
\]
and introduce the set
\[
D(N) : = \Big\{ y \in \hat{\mathcal{S}} : \frac{\mu_{y,M}(R(y))} {\nu_{y,M}(R(y))} \leq N \Big\}.
\]
Then for sufficiently large $N$, $m(D(N)) > 0$. The map
$ D(N)\ni y \mapsto h(y) : = \nu_{y,M}(R(y))/ \mu_{y,M}(R(y)) \leq N$ permits to define
\[
\hat \mu: = \int_{D(N)} h(y) \mu_{y,M} m (dy), \qquad \hat \nu : = \nu\llcorner_{R(D(N))\cap K_{2,M} }.
\]
It follows that $\hat \mu$ and $\hat \nu$ have bounded densities w.r.t. $\gamma$ and the set
$\hat{\mathcal{T}}: = \mathcal{T} \cap G(K_{1,\delta}) \cap G^{-1}(K_{2,\delta})$ is
a transport set for the transport problem between $\hat \mu$ and $\hat \nu$.
It follows from Theorem \ref{T:approx} and Theorem \ref{T:stimapprox} that $\hat \gamma:=\gamma\llcorner_{\hat {\mathcal{T}}}$
verifies Assumption \ref{A:NDEatom} w.r.t. $G \cap K_{1,M}\times X \cap X \times K_{2,M}$.
Therefore from Proposition \ref{P:nonatoms} follows that the conditional probabilities
$\hat \mu_{y}$ of the disintegration of $\hat \mu$ are continuous. Since $\hat \mu_{y} =c(y) \mu_{y}\llcorner_{\hat{\mathcal{T}}}$
for some positive constant $c(y)$, we have a contradiction.
\end{proof}
It follows straightforwardly the existence of an optimal invertible transport map.
\begin{theorem}\label{T:esiste}
Let $\mu,\nu \in \mathcal{P}(X)$ absolute continuous w.r.t. $\gamma$ and assume that there exists $\pi \in \Pi(\mu,\nu)$ such that $\mathcal{I}(\pi)$ is finite.
Then there exists a solution for the Monge minimization problem
\[
\min_{T : T_{\sharp}\mu=\nu} \int_{X} \| x-T(x) \|_{H(\gamma)} \mu(dx).
\]
Moreover we can find $T$ invertible.
\end{theorem}
\begin{proof}
For $m$-a.e $y \in \mathcal{S}$ $\mu_{y}$ and $\nu_{y}$ are continuous. Since $R(y)$ is one dimensional and
the ray map $\ni \mathbb{R} t \mapsto g(t,y)$ is an isometry w.r.t. $\| \cdot \|_{H(\gamma)}$, we can define
the non atomic measures $g(y,\cdot)_{\sharp}\mu_{y},g(y,\cdot)_{\sharp}\nu_{y} \in \mathcal{P}(\mathbb{R})$. By the one-dimensional theory, there exists
a monotone map $T_{y}:\mathbb{R} \to \mathbb{R}$ such that
\[
T_{y\,\sharp} \Big(g(y,\cdot)_{\sharp}\mu_{y} \Big) = g(y,\cdot)_{\sharp}\nu_{y}.
\]
Using the inverse of the ray map, we can define $T_{y}$ on $R(y)$. Hence for $m$-a.e. $y \in \mathcal{S}$ we have a $\| \cdot \|_{H(\gamma)}$-cyclically monotone map $T_{y}$ such that $T_{y\,\sharp}\mu_{y}=\nu_{y}$.
To conclude define $T : \mathcal{T} \to \mathcal{T}$ such that $T = T_{y}$ on $R(y)$. Indeed $T$ is $\mu$-measurable, invertible and $T_{\sharp}\mu=\nu$.
For the details, see the proof of Theorem 6.2 of \cite{biacava:streconv}.
\end{proof}
\section{Notation}
\label{S:notation}
\begin{tabbing}
\hspace{4cm}\=\kill
$P_{i_1\dots i_I}$ \> projection of $x \in \Pi_{k=1,\dots,K} X_k$ into its $(i_1,\dots,i_I)$ coordinates, keeping order
\\
$\mathcal{P}(X)$ or $\mathcal{P}(X,\Omega)$ \> probability measures on a measurable space $(X,\Omega)$
\\
$\mathcal{M}(X)$ or $\mathcal{M}(X,\Omega)$ \> signed measures on a measurable space $(X,\Omega)$
\\
$f \llcorner_A$ \> the restriction of the function $f$ to $A$
\\
$\mu \llcorner_A$ \> the restriction of the measure $\mu$ to the $\sigma$-algebra $A \cap \Sigma$
\\
$\mathcal{L}^d$ \> Lebesgue measure on $\mathbb{R}^d$
\\
$\mathcal{H}^k$ \> $k$-dimensional Hausdorff measure
\\
$\Pi(\mu_1,\dots,\mu_I)$ \> $\pi \in \mathcal{P}(\Pi_{i=1}^I X_i, \otimes_{i=1}^I \Sigma_i)$ with marginals $(P_i)_\sharp \pi = \mu_i \in \mathcal{P}(X_i)$
\\
$\mathcal{I}(\pi)$ \> cost functional \eqref{E:Ifunct}
\\
$c$ \> cost function $ : X \times Y \mapsto [0,+\infty]$
\\
$\mathcal{I}$ \> transportation cost \eqref{E:Ifunct}
\\
$\phi^c$ \> $c$-transform of a function $\phi$ \eqref{E:ctransf}
\\
$\partial^c \varphi$ \> $d$-subdifferential of $\varphi$ \eqref{E:csudiff}
\\
$\Phi_c$ \> subset of $L^1(\mu) \times L^1(\nu)$ defined in \eqref{E:Phicset}
\\
$J(\phi,\psi)$ \> functional defined in \eqref{E:Jfunct}
\\
$C_b$ or $C_b(X,\mathbb{R})$ \> continuous bounded functions on a topological space $X$
\\
$(X,d)$ \> Polish space
\\
$(X,d_L)$ \> non-branching geodesic separable metric space
\\
$D_L(x)$ \> the set $\{y : d_L(x,y) < +\infty\}$
\\
$\gamma_{[x,y]}(t)$ \> geodesics $\gamma : [0,1] \to X$ such that $\gamma(0) = x$, $\gamma(1) = y$
\\
$B_r(x)$ \> open ball of center $x$ and radius $r$ in $(X,d)$
\\
$B_{r,L}(x)$ \> open ball of center $x$ and radius $r$ in $(X,d_L)$
\\
$\mathcal{K}(X)$ \> space of compact subsets of $X$
\\
$d_H(A,B)$ \> Hausdorff distance of $A$, $B$ w.r.t. the distance $d$
\\
$L(X^{*},X)$ \> space of continuous and linear maps from $X^{*}$ to $X$
\\
$A_x$, $A^y$ \> $x$, $y$ section of $A \subset X \times Y$ \eqref{E:sectionxx}
\\
$\mathcal{B}$, $\mathcal{B}(X)$ \> Borel $\sigma$-algebra of $X$ Polish
\\
$\Sigma^1_1$, $\Sigma^1_1(X)$ \> the pointclass of analytic subsets of Polish space $X$, i.e.~projection of Borel sets
\\
$\Pi^1_1$ \> the pointclass of coanalytic sets, i.e.~complementary of $\Sigma^1_1$
\\
$\Sigma^1_n$, $\Pi^1_n$ \> the pointclass of projections of $\Pi^1_{n-1}$-sets, its complementary
\\
$\Delta^1_n$ \> the ambiguous class $\Sigma^1_n \cap \Pi^1_n$
\\
$\mathcal{A}$ \> $\sigma$-algebra generated by $\Sigma^{1}_{1}$
\\
$\mathcal{A}$-function \> $f : X \to \mathbb{R}$ such that $f^{-1}((t,+\infty])$ belongs to $\mathcal A$
\\
$h_\sharp \mu$ \> push forward of the measure $\mu$ through $h$, $h_\sharp \mu(A) = \mu(h^{-1}(A))$
\\
$\textrm{graph}(F)$ \> graph of a multifunction $F$ \eqref{E:graphF}
\\
$F^{-1}$ \> inverse image of multifunction $F$ \eqref{E:inverseF}
\\
$F_x$, $F^y$ \> sections of the multifunction $F$ \eqref{E:sectionxx}
\\
$\mathrm{Lip}_1(X)$ \> Lipschitz functions with Lipschitz constant $1$
\\
$\Gamma'$ \> transport set \eqref{E:gGamma}
\\
$G$, $G^{-1}$ \> outgoing, incoming transport ray, Definition \ref{D:Gray}
\\
$R$ \> set of transport rays \eqref{E:Rray}
\\
$\mathcal{T}$, $\mathcal{T}_e$ \> transport sets \eqref{E:TR0}
\\
$a,b : \mathcal{T}_e \to \mathcal{T}_e$ \> endpoint maps \eqref{E:endpoint0}
\\
$\mathcal S$ \> cross-section of $R \llcorner_{\mathcal T \times \mathcal T}$
\\
$g = g^+ \cup g^-$ \> ray map, Definition \ref{D:mongemap}
\\
$Jac(T)(x)$ \> Jacobian determinant $|\det(dT(x))|$, Theorem \ref{T:stimapprox}
\end{tabbing}
|
\section{Introduction}\label{Sect:introduction}
Cosmic structures are supposed to rise from primordial matter fluctuations generated at very early times, during the inflation era.
These fluctuations represent the seeds that grew over the cosmic time till the formation of the presently observed Universe \cite[e.g.][]{Komatsu2011}.
In the standard scenario, a Gaussian distribution of such perturbations is assumed, as a consequence of the central limit theorem.
\\
However, cosmic microwave background (CMB) temperature fluctuations, at high-order perturbation theory \cite[see e.g.][]{Zaldarriaga2000,Bartolo2004,Seery2005,Assadullahi2007,Cooray2008,Beltran2008,Liguori2008,Khatri2009,Lam2009,Bartolo2010,Yadav2010,Gao2010,Pitrou2010}, could have deviations from the purely Gaussian shape and leave room for non-Gaussian models.
Additional studies of the excursion set formalism and analyses of N-body numerical simulations of large-scale structures have shown how the abundance of rare-peak dark-matter haloes and their clustering properties and distribution might be affected \cite[e.g.][]{Grinstein1986,Koyama1999,Robinson2000,Grossi2007,Kang2007,Dalal2008,Grossi2009,Desjacques2009,Lam2010,Maggiore2010,Pillepich2010,Roncarelli2010,Pace2010,Wagner2010,LoVerde2011arXiv,Maturi2011arXiv,Yokoyama2011arXiv}.
\\
Observationally, the CMB is the best way to probe non-Gaussianities and different experiments (COBE, BOOMERanG, WMAP, PLANCK) can be used \cite[e.g.][]{Komatsu2002,Komatsu2003,Gaztanaga2003,Spergel2007,Hikage2008,Yadav2008,Afshordi2008,Slosar2008,Komatsu2009,Natoli2009,Komatsu2010,Hou2010,Raeth2009,Raeth2010arXiv}.
Recent determinations of the cosmological parameters by the 7-year Wilkinson Microwave Anysotropy Probe (WMAP) satellite \cite[e.g.][]{Komatsu2011} still find primordial non-Gaussianities.\\
Despite the huge amount of work based on N-body dark-matter only simulations, detailed chemical and hydrodynamical investigations of the effects on the visible matter are still missing, or poorly understood.
According to previous numerical studies \cite[][]{Viel2009}, the Lyman-$\alpha$ flux bispectrum in the high-redshift intergalactic medium is expected to present deviations from the Gaussian case up to $\sim 10$ per cent at $z\sim 4$.
Semianalytic estimates \cite[e.g.][]{Crociani2009} seem to show that the ionized fraction of the intergalactic medium (IGM) can grow by a factor of a few, up to 5, with respect to the corresponding Gaussian model.
The increase of the filling factor has a small impact on the reionization optical depth and is of the order of $\sim 10$ per cent if a scale-dependent non-Gaussianity is assumed.
Since non-Gaussianities are expected to influence the primordial matter density fluctuations, it is important to address their consequences on the history of the baryons, from the early ``dark ages'', when stars were not formed, yet, to the following ``louminous ages'', after the birth of the first objects.
Additionally, it is relevant to understand the role they play in the transition from the primordial population III (popIII) stars to present-day population II-I (popII) stars.
These are indeed among the goals of the present work.
\\
A standard way to account for the existence of non-Gaussianities in the primordial fluctuation field is to regard them as a perturbation on a Gaussian background. Strongly non-Gaussian models have been progressively dismissed by CMB observations and it is now generally agreed that primordial non-Gaussianities, if present, must have represented a moderate deviation from the underlying Gaussian statistics. The most commonly adopted parametrisation of non-Gaussianities considers them as second-order perturbations of the Bardeen gauge-invariant potential \cite[see e.g.][]{Salopek1990,Komatsu2001,Verde2010,Desjacques2010}:
\begin{equation}\label{eq:nong}
\Phi = \Phi_{\rm L} + f_{\rm NL} \left[ \Phi_{\rm L}^2 - <\Phi_{\rm L}^2> \right],
\end{equation}
where $\Phi_{\rm L}$ is the {\it linear} Gaussian part, and the dimensionless coupling constant, \fnl, controls the importance of the non-Gaussian contribution\footnote{
The Bardeen gauge-invariant potential, $\Phi$, reduces to the usual Newtonian peculiar gravitational potential, up to a minus sign, on sub-Hubble scales.
In the literature \cite[e.g.][]{Verde2010}, there are two, sometimes misleading, notations: the large-scale structure (LSS) convention and the CMB convention.
In the former (LSS), $\Phi$ is linearly extrapolated at $z=0$;
in the latter (CMB), $\Phi$ denotes the primordial potential, at $z\rightarrow +\infty$.
The net difference is related to the ratio of the growth factor, $g(z)$, at the two times and accounts for a correction of roughly unity:
i.e.
\fnl$^{\rm LSS}$
$= g(z\rightarrow +\infty)/g(z=0)\, $
\fnl$^{\rm CMB}$,
which corresponds to $ \sim 1.28$\fnl$^{\rm CMB}$, for a standard $\Lambda$CDM model,
and to $ \sim 1.33$\fnl$^{\rm CMB}$, for the latest observational data from WMAP \cite[][]{Komatsu2011}.
Usually, observational values are reported according to the CMB convention.
}.
Since the final value of the potential $\Phi$ depends on the local value of the Gaussian field $\Phi_{\rm L}$, non-Gaussianities of the the form (\ref{eq:nong}) are labelled as ``local''. Local models are physically relevant for scenarios where non-linearities develop outside the horizon, but they are by no means exhaustive of all the forms non-Gaussianities can assume according to the several mechanisms that generate them \cite[for further details see e.g][and references therein]{Babich2004,Bartolo2004,Loverde2008}.
The exact value of the \fnl{} parameter is continuously better constrained by CMB observations (see discussion in Sect.~\ref{Sect:discussion}).
More extended analyses could in principle be done by introducing additional parameters, like the cubic-order one, \gnl, but, given their large uncertainties \cite[e.g. $-7.4\times 10^5<$\gnl$<8.2\times 10^5$, according to][]{Smidt2010}, we will consider only second-order corrections.
In addition, since latest determinations suggest \fnl$\gtrsim 0$ \cite[][]{Spergel2007,Slosar2008,Afshordi2008,Yadav2008,Curto2009b,Komatsu2009,Smith2009,Rudjord2009,Komatsu2010,Hou2010,Rudjord2010,Cabella2010}, we will only consider positive values for \fnl.
\\
In the present work we will focus on the effects non-Gaussianities have on the baryon history in universes with different initial conditions, by studying how gas evolution, cooling and condensation could lead to different star formation epochs for different initial scenarios, and by exploring how the various stellar population regimes and the statistical properties of gas and stars are affected in non-Gaussian models.\\
We will present the simulations performed in Sect.~\ref{Sect:simulations},
together with the main results, in Sect.~\ref{Sect:results}, about
the first cosmic structures (Sect.~\ref{Sect:maps}), star formation (Sect.~\ref{Sect:SFR}), chemistry evolution (Sect.~\ref{Sect:chemistry}) and primordial streaming motions \cite[see also e.g.][]{TseliakhovichHirata2010,Maio2011}, and baryon statistics (Sect.~\ref{Sect:distributions}).
Then, we will discuss and conclude in Sect.~\ref{Sect:discussion}.
\section{Simulations}\label{Sect:simulations}
\begin{table*}
\centering
\caption[Simulation set-up]{Initial parameters for the different runs.}
\begin{tabular}{lccccccc}
\hline
\hline
Runs& Box side & Particle mass [$\msun/h$] for & Softening &\fnl & Gas bulk & PopIII IMF\\
& [\Mpch] & gas (dark matter) & [kpc/{\it h}]& & velocity [km/s] & range [M$_\odot$]\\
\hline
Run05.0 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 0 & 0 & [100, 500]\\
Run05.10 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 10 & 0 & [100, 500]\\
Run05.50 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 50 & 0 & [100, 500]\\
Run05.100 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 100 & 0 & [100, 500]\\
Run05.1000 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 1000 & 0 & [100, 500]\\
\hline
Run100.0 & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 0 & 0 & [100, 500]\\
Run100.10 & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 10 & 0 & [100, 500]\\
Run100.50 & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 50 & 0 & [100, 500]\\
Run100.100 & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 100 & 0 & [100, 500]\\
Run100.1000 & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 1000 & 0 & [100, 500]\\
\hline
Run100.0.SL & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 0 & 0 & [0.1, 100]\\
Run100.10.SL & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 10 & 0 & [0.1, 100]\\
Run100.50.SL & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 50 & 0 & [0.1, 100]\\
Run100.100.SL & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 100 & 0 & [0.1, 100]\\
Run100.1000.SL & 100 & $3.39\times 10^8\quad$ ($2.20\times 10^9$)& 7.8 & 1000 & 0 & [0.1, 100]\\
\hline
Run05.0.30 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 0 & 30 & [100, 500]\\
Run05.0.60 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 0 & 60 & [100, 500]\\
Run05.0.90 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 0 & 90 & [100, 500]\\
Run05.100.30 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 100 & 30 & [100, 500]\\
Run05.100.60 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 100 & 60 & [100, 500]\\
Run05.100.90 & 0.5 & $42.35\quad$ ($275.28$)& 0.04 & 100 & 90 & [100, 500]\\
\hline
\label{tab:runs}
\end{tabular}
\begin{flushleft}
\vspace{-0.5cm}
{\small
}
\end{flushleft}
\end{table*}
We use a modified version of the parallel tree/SPH numerical code P-Gadget2 \cite[][]{Springel2005}.
Beyond gravity and hydrodynamics, it includes radiative gas cooling at high \cite[][]{SD1993} and low \cite[][]{Maio2007} temperatures, both from molecules and fine-structure metal transitions, multiphase model \cite[][]{Springel2003} for star formation \cite[inspired on the works by][]{KatzGunn1991,Cen1992,CenOstriker1992,Katz1992,Katz_et_al_1992,Katz_et_al_1996}, UV background radiation \cite[][]{HaardtMadau1996}, and wind feedback \cite[][]{Springel2003,Aguirre_et_al_2001}.
\\
The main agents of gas cooling, in hot environments, are atomic resonant transitions of H and He excitations that can rapidly bring the temperatures down to $\sim 10^4\,\rm K$.
In metal-rich environments, atomic fine-structure transitions can further lower them \cite[see][]{Maio2007}, mostly at high redshift, since the formation of a UV background \cite[][]{HaardtMadau1996,Haehnelt2001,Gilmore2009} at lower redshifts competes by heating the IGM.
In dense environments, stochastic star formation is assumed, cold gas is gradually converted into stars (that reproduce the Kennicut-Schmidt low), outflows take place, and feedback effects inject entropy into the surrounding environment, leading to a self-regulated star forming regime \cite[for further detail see ][]{Springel2003}.
\\
In the code, we also include the relevant chemical network to self-consistently follow the cosmic evolution of e$^-$, H, H$^+$, H$^-$, He, He$^+$, He$^{++}$, H$_2$, H$_2^+$, D, D$^+$, HD, HeH$^+$ \cite[see e.g.][ and references therein]{Yoshida2003,Maio2007,Maio2009,Maio2010a}, metal (C, O, Si, Fe, Mg, S) pollution from popIII and/or popII stellar generations, ruled by a critical metallicity threshold of $Z_{crit}=10^{-4}\,\zsun$ \cite[][]{Tornatore2007,Maio2010a} and leading gravitational enrichment into the surrounding medium \cite[][]{Maio2011arXiv}.
\\
We perform several runs (see Tab.~\ref{tab:runs}) in order to check the different assumptions on \fnl, at large and small scales.
\\
The initial conditions are generated with a modified version of the N-GenIC code that introduces non-Gaussian features of the local type in the initial density field according to eq. (\ref{eq:nong}) \cite[for further details see e.g.][]{Bartolo2005,Grossi2007,Grossi2009,Viel2009}.
We assume a concordance $\Lambda$CDM model with
matter density parameter $\Omega_{\rm 0,m}=0.3$,
cosmological density parameter $\Omega_{\rm 0,\Lambda}=0.7$,
baryon density parameter $\Omega_{\rm 0,b}=0.04$,
expansion rate at the present of H$_0=70\,\rm km/s/Mpc$,
power spectrum normalization via mass variance within 8~Mpc/{\it h} radius sphere $\sigma_8=0.9$,
and spectral index $n=1$.
We study the non-Gaussianity parameter space identified by \fnl = 0, 10, 50, 100, 1000, in cosmological periodic boxes of 0.5~\Mpch{ } and 100~\Mpch{} a side, sampled, at redshift $z=100$, with $2\times 320$ particles with gas mass resolution of $\sim 40\,\msunh$ and $\sim 3\times 10^8\,\msunh$, and corresponding maximum physical softening lengths of 0.04~\kpch{} and 7.8~\kpch, respectively.
This will allow us to see the impacts of the different models down to scales of the order of $\sim 0.1\,\rm kpc$.
We note that the small size\footnote{
Results on biased high-density (zoomed) initial conditions, for which the onset of star formation is expected to take place at redshifts as high as $\sim 40-50$ heve been presented by e.g. \cite{Maio2009}.
While in such set-ups the timing of the popIII formation might be more accurate and independent from Lyman-Werner radiation feedback \cite[][]{Johnson2008,TrentiStiavelli2009}, the general physical trends are supposed to be very similar.
}
of the 0.5~\Mpch{} boxes prevents the simulations from being run to very low redshifts, in order for the fundamental mode to remain linear; the large simulations satisfy instead this requirement \cite[see e.g. Sect.~2.3 in][]{Maio2009PhDT} and are thus run until $z=0$.
The simulations will be labelled as Run05.0, Run05.10, Run05.50,Run05.100, Run05.1000, and Run100.0, Run100.10, Run100.50,Run100.100, Run100.1000, respectively.\\
Throughout this work, we will assume a Salpeter IMF for popII star forming regions (where $Z\ge Z_{crit}$), and a top-heavy IMF, with mass range [100$\msun$, 500$\msun$], for popIII star forming regions (where $Z<Z_{crit}$).
The latter assumption is matter of debate, since there are also theoretical and numerical evidences for the existence of low-mass popIII stars \cite[e.g.][]{Yoshida2006, Yoshida_et_al_2007,CampbellLattanzio2008,SudaFujimoto2010,Stacy2010}.
If this is the case, that might mean that the conclusions of several previous works predicting massive primordial stars were highly affected by numerics and resolution.
As the primordial popIII IMF can impact the cosmological pollution history \cite[e.g.][]{Maio2011arXiv}, in Sect.~\ref{Sect:SFR}, we we will also consider, for the 100~\Mpch{ } side boxes, the opposite extreme of a Salpeter-like popIII distribution, with appropriate yields, for primordial stars \cite[see discussion in][and references therein]{Maio2010a}, and mass range [0.1$\msun$, 100$\msun$]: the corresponding runs will be labelled by adding the SL suffix.\\
In order to see how relevant the consequences of non-Gaussian perturbations are, in Sect.~\ref{Sect:SFR}, we will also compare them with second-order effects in the linearized perturbation theory \cite[][]{TseliakhovichHirata2010}, that expect supersonic gas bulk flows at early times to suppress structure growth, and delay star formation on scales smaller than a few Mpc \cite[][]{Maio2011}.
We will focus on the \fnl=0 and \fnl=100 cases, assuming primordial bulk shifts of 30, 60, 90~km/s.
The six corresponding simulations will be identified by Run05.0.30, Run05.0.60, Run05.0.90 (for the \fnl=0 set), and by Run05.100.30, Run05.100.60, Run05.100.90 (for the \fnl=100 set), respectively.\\
A friend-of-friend (FoF) algorithm \cite[][]{Springel2001}, with comoving linking lenght of 20 per cent the mean inter-particle separation, is applied at postprocessing time, considering all the types of particles, to find the formed cosmic objects, with their dark, gaseous, and stellar components.\\
We summarize the properties of all the different runs in Table~\ref{tab:runs}.
\section{Results}\label{Sect:results}
In the following, we will show the main repercussions primordial non-Gaussianities have on the cosmic baryon history in the Universe.
We will focus on cosmic structure evolution (Sect.~\ref{Sect:maps}), star formation (Sect.~\ref{Sect:SFR}), chemical properties of the Universe (Sect.~\ref{Sect:chemistry}), and baryon statistics (Sect.~\ref{Sect:distributions}) in different models.
\begin{figure*}
\centering
\includegraphics[width=\textwidth]{./Figure/Maps/mapsT_all.eps}
\caption[Maps]{\small
Mass-weighted temperature evolution for the large-scale 100~\Mpch{} side boxes, with \fnl=0 (first and second row), and \fnl=1000 (third and fourh row), at different redshifts (see legends).
All the maps refer to slices centered at the mid-plane of the box, height $z=50\,\rm\Mpch$, and have thickness of $\sim 7\,\rm\Mpch$, i.e. $1/14$ the boxsize.
}
\label{fig:Maps}
\end{figure*}
\subsection{Structure evolution}\label{Sect:maps}
We start by showing the basic features of the simulations.
In Fig.~\ref{fig:Maps}, we give a pictorial representation of the simulation with side of 100~\Mpch, and display the mass-weighted temperature evolution for a slice centered in the middle of the box, along the z-direction, having a thickness of $\sim 7\,\rm\Mpch$, and smoothed over a grid of 1024$\times$1024 pixels.
We consider the \fnl=0 (first two rows) and the \fnl=1000 (last two rows) cases, and highlight the differences among them.
At high redshift, $z\gtrsim 20$, the \fnl=1000 case shows more perturbations with respect to the \fnl=0 one, and this is visible at $z\sim 10-15$, as well, when feedback effects from early star formation heat the medium above $\sim 10^4-10^5\,\rm K$.
The maps of the \fnl=1000 case present more advanced stages of structure formation, and this keeps the gas slightly hotter, with temperatures of $\sim 2\times 10^6\,\rm K$, until $z\sim 6$, i.e. for about the first Gyr of the Universe.
Temperature variations are up to a factor of $\lesssim 2$ for the \fnl=1000 case, but only up to $\sim 10$ per cent, for the other ones.
Later on, at $z<6$, more differences are not clearly visible and the final trends at lower redshift ($z\sim 1$) catch up and lead to the same behaviour at $z\sim 0$.\\
We conclude that baryon history is affected by primordial non-Gaussianities mostly at early times, and this statement is strengthened from Fig.~\ref{fig:Maps2}, where we plot the molecular fraction at redshift $z=25$, for the high-resolution runs, with side of 0.5~\Mpch.
We compare the \fnl=0, the \fnl=100, and the \fnl=1000 cases during the very first stages of structure formation via molecular (mainly H$_2$ and HD) cooling and catastrophyc run-away collapse.
While in low-density environments there are no strong net differences, in the high-density regions the molecular content can change up to a factor of a few.
Indeed, in the \fnl=100 case, molecules increase of $\sim 10$ per cent (more exactly of 6 per cent) with respect to the \fnl=0 one, and in the \fnl=1000 case, of a factor of $\sim 2.5$.
\\
In Fig.~\ref{fig:Maps3}, we highlight the differences by plotting the contrast between the \fnl=100 and \fnl= 0 cases (top), and between the \fnl=1000 and \fnl=0 cases (bottom), defined, for each pixel in the map, as
\begin{equation}
C_{\rm molecules} = \frac{x_{\rm molecules, 1}-x_{\rm molecules, 0}}{x_{\rm molecules, 1}+x_{\rm molecules, 0}}.
\end{equation}
In the previous equation, $x_{\rm molecules, 1}$ is the molecular fraction in the pixel for the \fnl=100 or the \fnl=1000 case; while $x_{\rm molecules, 0}$ is the reference \fnl=0 case.
It is clear that the comparison between \fnl=100 and \fnl=0 expects much smaller differences than the one between \fnl=1000 and \fnl=0.
Indeed, the top panel presents about 54 per cent of the entire area corresponding to contrasts of $\lesssim 10^{-3}$ (blue), almost 45 per cent corresponding to contrasts of $\sim 10^{-3}-10^{-2}$ (cyan-green), and the remaining few per cent reaching values of $\sim 10^{-1}$, or larger (yellow-red).
On the other side, the bottom panel displays larger deviations from the reference \fnl=0 case, with only 15 per cent of the region having contrasts $\lesssim 10^{-3}$ (blue), roughly 70 per cent of $\sim 10^{-3}-10^{-2}$ (cyan-green), and the remaining 15 per cent larger values (yellow-red).
This means that larger \fnl{} will induce higher molecular production, stronger cooling, earlier gas collapse and earlier star formation, as we will see in the next section.
\\
The fundamental reason why gas and baryonic structures are sensitive to different \fnl{} is that primordial non-Gaussianities affect the growth and evolution of the underlying dark matter haloes.
As a consequence, the resulting baryon history reflects the imprints of the primordial dark-matter non-Gaussian distribution.
\begin{figure*}
\centering
\includegraphics[width=\textwidth]{./Figure/Maps/mapsX_all.eps}
\\
\caption[Maps2]{\small
Molecule maps at redshift $z=25$ for the high-resolution 0.5~\Mpch side boxes, with \fnl=0 (left), \fnl=100 (center), and \fnl=1000 (right).
All the maps refer to slices centered at the mid-plane of the box, height $z=0.25\,\rm\Mpch$, and have thickness of $\sim 0.03\,\rm\Mpch$, i.e. $1/14$ the boxsize.
}
\label{fig:Maps2}
\end{figure*}
\begin{figure}
\centering
\includegraphics[width=0.33\textwidth]{./Figure/Maps/0.5Mpch/fnl100/MOLECULES_difference_008.eps}\\
\fbox{\fnl=100 -- \fnl=0}\\
\includegraphics[width=0.33\textwidth]{./Figure/Maps/0.5Mpch/fnl1000/MOLECULES_difference_008.eps}\\
\fbox{\fnl=1000 -- \fnl=0}\\
\caption[Maps3]{\small
Molecular fraction contrast at redshift $z=25$ between the \fnl=100 and \fnl=0 cases (top), and between the \fnl=1000 and \fnl=0 cases (bottom).
The color scales refer to the logarithm of the absolute value of the molecular fraction contrasts in the two cases.
}
\label{fig:Maps3}
\end{figure}
\subsection{Star formation}\label{Sect:SFR}
\begin{figure*}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/SFR/sfr_ratio_average_Zcrit_4_nonG_05Mpch.ps}
\includegraphics[width=0.45\textwidth]{./Figure/SFR/sfr_ratio_average_Zcrit_4_nonG_100Mpch.ps}\\
\caption[SFR]{\small
Star formation rate densities as a function of the redshift for the simulation runs of the 0.5~\Mpch{} side boxes (left) and 100~\Mpch{} side boxes (right).
In the top panels, the upper lines refer to the total star formation rate densities, while the bottom lines to the popIII ones.
In all the cases, we consider
\fnl=0 (black solid lines),
\fnl=10 (cyan dotted lines),
\fnl=50 (green short-dashed lines),
\fnl=100 (red long-dashed lines), and
\fnl=1000 (blue dotted short-dashed lines.
They correspond to Run05.0, Run05.10, Run05.50, Run05.100, Run05.1000, on the left panels,
and to Run100.0, Run100.10, Run100.50, Run100.100, Run100.1000, on the right panels.
Below, we plot the corresponding popIII contributions to the total star formation rates.
}
\label{fig:SFR}
\end{figure*}
At this point, we study the star formation history in the different models, paying attention to both the popIII and the popII regime.\\
In Fig.~\ref{fig:SFR}, we plot the star formation rate densities from the simulations of the 0.5~\Mpch{} side boxes (left), and 100~\Mpch{} side boxes (right), with the relative popIII contribution (lower panels).
The simulations have similar trends, but the onset of star formation is slightly earlier for the runs in 0.5~\Mpch{} side boxes.
This is essentially due to the fact that the higher-resolution simulations are able to capture better molecular cooling and gas collapse in primordial mini-haloes.
Thus, they can address star formation in those regimes (i.e. at masses smaller than $\sim 10^{8}\msunh$) where the larger 100~\Mpch{} side boxes are limited by resolution (see properties in Tab.~\ref{tab:runs}).
We note that at both small and large scales, the main effect of non-Gaussianities is a general shift of star formation, for the two population regimes.
This is due to the fact that at larger \fnl{} the contribution to the exponential tail of the mass function increases, leading to higher and earlier star formation.
The differences among the various onsets are very small for \fnl=0, \fnl=10, \fnl=50 and \fnl=100, and correspond to some $\sim 10^7\,\rm yr$.
Only for the \fnl=1000 case there is a larger gap of $\Delta z\sim$ a few, i.e. $\sim 5\times 10^7\,\rm yr$, in the 0.5~\Mpch{} box, and $\sim 10^8\,\rm yr$, in the 100~\Mpch{} box.
The behaviours of the popIII regimes are very little affected and their overall contributions drop down to $\sim 10^{-3} - 10^{-2}$, for any \fnl, by redshift $z\sim 10$.
Due to the anticipated evolution of the runs with larger \fnl, the corresponding popIII contributions drop down earlier and stay slightly below the \fnl=0 cases.
On small scales (left panels), the difference is about $\sim 2$ orders of magnitude, at $z\sim 19-20$, but only a factor of a few at later stages.
At $z\sim 10$, in the \fnl=0 case one expects a popIII contribution of $\sim 0.0010$, while in the \fnl=1000 case, of $\sim 0.0008$ -- 20 per cent smaller.
Similarly, at the same redshift, on larger scales (righ panels), the differences are well within a factor of 2, with a popIII contribution of $\sim 0.0097$ for \fnl=0, and $\sim 0.0055$ for \fnl=1000.
\\
To check for degeneracies among different physical processes involved in the description of cosmic structure evolution, we have studied also how the scenario changes in two additional cases:
\begin{itemize}
\item
by adopting a common Salpeter-like IMF for the primordial popIII generation;
\item
by comparing with second-order corrections to the linear perturbation theory \cite[][]{TseliakhovichHirata2010}, which are supposed to generate primordial gas bulk motions, to suppress star formation on scales smaller than a few Mpc \cite[][]{Maio2011}, and to induce delays in the reionization of the Universe \cite[][]{Dalal2010}.
\end{itemize}
In Fig.~\ref{fig:SFR_IMFrange}, we plot the star formation rate densities for the 100~\Mpch{} side boxes with primordial \fnl=0, \fnl=10, \fnl=50, \fnl=100, and \fnl=1000.
In these cases (see details in Tab.~\ref{tab:runs}), we have considered a Salpeter-like popIII IMF that is expected to cause a longer popIII-dominated epoch, because of the loger lifetimes of the $\sim 10\rm s\,\msun$ stars dying as supernovae \cite[see][]{Maio2010a}.
In fact, we find that the popIII regime dominates at early times and its contribution drops down to $\sim 10$ per cent only after $\sim 4\times 10^8\,\rm yr$.
Furthermore, the global popIII contributions are higher in comparison with the top-heavy popIII IMF previously discussed, and the difference is more than one order of magnitude.
At redshift $z\sim 10$, it is roughly a few times $10^{-1}$, which means that the primordial IMF can have impacts on the baryon and stellar evolution that are much stronger than the non-Gaussian effects.
Therefore, if we want to give constraints on primordial non-Gaussianities it will be important first to properly constrain the popIII IMF.\\
About the second check, we use only the small-box simulations, since they are more suitable to take into account primordial streaming motions.
These are predicted to have rms velocities of the order of $\sim 30\,\rm km/s$ at decoupling \cite[][]{TseliakhovichHirata2010}.
So, we focus (see details in Tab.~\ref{tab:runs}) on the same initial conditions described above for \fnl=0 and \fnl=100, and add \cite[as in][]{Maio2011} a primordial bulk shift of 30~km/s to the gas.
To consider statistical deviations from the rms value, we additionally run the cases with 60~km/s and 90~km/s shifts (i.e. 2 and 3 times the rms value), as upper limits, which can be found in rare regions (less than $\sim 1$ per cent) of the Universe, though.
In Fig.~\ref{fig:SFR_vb}, we compare the results for the two different \fnl, and plot the \fnl=0 and no-shift case (thin solid line) together with all the \fnl=100 cases (thick lines).
Obviously, star formation sets in at different times, according to the different values of the streaming motions and the resulting time delays are of the order of tens of Myr ($\Delta z\sim$ a few).
However, if we compare the \fnl=0 (thin solid line) and the \fnl=100 (thick lines) cases for a fixed bulk shift velocity, we notice that the effects of non-Gaussianities are comparable to or negligible with respect to the impacts of the streaming motions.
The star formation rate in the \fnl=100 and 30~km/s-shift (thick dotted line) case almost coincides with the one in the \fnl=0 run, and this leads to a degeneracy between these two different phenomena.
Moreover, in the \fnl=100 and 60~km/s-shift (thick short-dashed line) case and in the \fnl=100 and 90~km/s-shift (thick long-dashed line) case streaming motions cause much stronger delays.
The value of 30~km/s is the average expected in the whole Universe, but different values can be found in different regions, thus, purely non-Gaussian effects are going to be strongly twisted with the statistical variations of such second-order contaminations.
\\
Finally, we stress that variations on the determination of $\sigma_8$ of a few percents could produce a similar degeneracy, and different primordial Lyman-Werner (LW) radiation strengths coming from the early star forming episodes could probably have even larger effects.
In fact, the LW radiation (which is not explicitely followed in the simulations persented here) could dissociate primordial molecules, partially inhibiting H$_2$ cooling in the neighbouring sites.
Thus, the star formation process and the popIII formation at early times are expected to be delayed to lower redshifts.
But the metals which are then ejected from primordial stars are insensitive to LW radiation: independently from molecule formation, they strongly enhance the cooling capabilities of the surrounding gas at any temperature \cite[e.g.][]{Maio2007}, and can rapidly lead to a popII-dominated regime \cite[][]{Maio2010a,Maio2011}.
Given the tight connections between primordial IMF, LW radiation and chemical feedback more detailed studies will be needed to draw definitive conclusions on this issue, but definitely these will not change the results related to the \fnl{} effects.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/SFR/sfr_ratio_average_Zcrit_4_nonG_100Mpch_IMFrangeSalpeter.ps}
\caption[SFR Salpeter PopIII IMF]{\small
Star formation rate densities as a function of the redshift for the simulations of the 100~\Mpch{} side boxes, with Salpeter popIII IMF.
In the top panels, the upper lines refer to the total star formation rate densities, while the bottom lines to the popIII ones.
In all the cases, we consider
\fnl=0 (black solid lines),
\fnl=10 (cyan dotted lines),
\fnl=50 (green short-dashed lines),
\fnl=100 (red long-dashed lines), and
\fnl=1000 (blue dotted short-dashed lines.
They correspond to Run100.0.SL, Run100.10.SL, Run100.50.SL, Run100.100.SL, Run100.1000.SL.
Below, we plot the corresponding popIII contributions to the total star formation rates.
}
\label{fig:SFR_IMFrange}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/SFR/sfr_ratio_average_Zcrit_4_nonG_05Mpch_vb.ps}
\caption[vb]{\small
Upper panel: star formation rate densities as a function of the redshift for the simulation runs of the 0.5~\Mpch{} side boxes with \fnl=100, and primordial gas bulk velocities of
0~km/s (thick black solid line),
30~km/s (thick cyan dotted lines),
60~km/s (thick green short-dashed lines), and
90~km/s (thick red long-dashed lines).
As a comparison, we overplot the results for the 0.5~\Mpch{} side box with \fnl=0 and 0~km/s for the primordial gas bulk velocity (thin black solid line).
Bottom panel: the corresponding popIII contributions to the total star formation rates.
}
\label{fig:SFR_vb}
\end{figure}
\subsection{Chemical evolution}\label{Sect:chemistry}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/Zevolution/redshift_met_fnl_All_txt.ps}
\caption[Zevolution]{\small
Redshift evolution of the average pollution in the enriched regions (upper lines), and the average pollution in the whole box (lower lines), for \fnl=0 (solid lines), \fnl=100 (short-dashed lines), and \fnl=1000 (long-dashed lines), in the 100~\Mpch{} side boxes.
The horizontal dot-dot-dot-dashed line is drawn at the critical value of $10^{-4}\,Z_\odot$.
}
\label{fig:Zcompare}
\end{figure}
We display, in Fig.~\ref{fig:Zcompare}, the metallicity evolutions for the runs of the \fnl=0, \fnl=100, and \fnl=1000 cases, in the 100~\Mpch{} side box with a top-heavy popIII IMF.
The upper lines correspond to the average enrichment (i.e. the average metallicity of the polluted regions only), and the lower lines to the average metallicity (in the whole simulated volume), as a function of the redshift.
What is particularly evident is the rapidity of early metal pollution \cite[consistently with][]{Maio2010a}.
Indeed, the critical value, $Z_{crit}=10^{-4}\,\rm Z\odot$, is reached in only a few $10^7\,\rm yr$, from the first episodes of star formation and the primordial yields are so high that efficiently pollute the surrounding medium \cite[see more discussion in][]{Maio2011arXiv}.
The following metal evolution shows an increasing trend that leads to an average enrichment of about $Z\sim 10^{-3}-10^{-2}$, i.e. $\sim 0.1\,Z_\odot$.
Due to the patchyness of the metal enrichment process, it is possible to get very strong local pollution, too, with highly supersolar $Z$.
The average metallicity in the whole volume (solid line) increases more gradually, by changing of $\sim 6$ orders of magnitude between $z\sim 16$ and $z\sim 6$.
This behaviour is due to the fact that, as cosmic time evolves, the regions involved in metal pollution are larger and larger, in comparison to the very first isolated events.
However, the average enrichment reaches the critical value, $Z_{crit}$, by redshift $z\sim 6$, when the Universe is roughly one-billion-year old.
Afterwards, during the subsequent $\sim 13\,\rm Gyr$, $Z$ increases of $\sim 3$ orders of magnitude, and the cosmic star formation history is fully dominated by the popII regime (see also Fig.~\ref{fig:SFR}).
Among the different elements, oxygen is the dominant one, since it is heavily produced by PISN and SNII explosions.
While at early times it dominates by about one order of magnitude, at later stages other elements become important, in particular, iron, which is mostly produced by the death of long-living stars.
It catches up carbon and overtakes $\alpha$-elements, like magnesium, sulphur, and silicon.
\\
Briefly speaking, the fact that the average enrichment reaches the critical value already by $z\sim 6$ has broad implications also for reionization, given that metallicity affects the spectral energy distributions (SEDs) of stellar populations \cite[e.g.][]{Schaerer2003}.
In particular, primordial stellar populations have UV fluxes (at wavelengths shorter than $912\,\rm\AA$) which are up to four times larger than the corresponding popII ones, and in principle, could ionize hydrogen in a much more efficient way.
However, given the high level of enrichment in the clustered star forming regions (upper lines in Fig.~\ref{fig:Zcompare}) the popIII star formation rate at $z\sim 6$ is only roughly $\sim 10^{-2}-10^{-4}$ the total one (depending on the modeling).
The relative number fraction of popIII hundred-solar-masses stars with respect to the corresponding popII stars drops further down of a factor of $\sim 10-100$.
As a consequence, it is unlikely that popIII stars will dominate the reionization process simply because they are too rare and, additionally, shine for a too short time (up to few $10^6\,\rm yr$) to provide significant amounts of UV photons.
\\
These conclusions confirm and extend to lower redshift the finding by \cite{Maio2010a}, and show how primordial non-Gaussianities affect the whole picture.
In fact, by comparing the different \fnl{} cases, it emerges that the strongest difference is found for the \fnl=1000 run, in which metal enrichment kicks in at earlier times, tracing back the star formation rate (see Fig.~\ref{fig:SFR}).
Its behaviour is not drammatically different from the other ones, below $z \lesssim 12$, though.
We note that the \fnl=0 and \fnl=100 runs predict extremely similar evolutions, with average metallicities differing by a factor of a few only at the very beginning of the pollution process (at $z\sim 13-15$, where the average enrichment is of $Z<10^{10}$), and rapidly converging later on.
\\
The trends found in all the other runs show very similar behaviour and lead to the same conclusions.
\subsection{Baryon distributions}\label{Sect:distributions}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/Ngas/check_mfct_fnl_gasonly_compare_cumulative008.ps}\\
\includegraphics[width=0.45\textwidth]{./Figure/Ngas/check_mfct_fnl_gasonly_compare_cumulative018.ps}\\
\caption[Ngas]{\small
Cumulative primordial-cloud distributions as a function of gas mass, for $z=25$ (upper panel) and $z=15$ (lower panel), for the runs with \fnl=0 (black solid lines), \fnl=100 (red long-dashed lines), and \fnl=1000 (blue dotted short-dashed lines).
We plot results for the high-resolution simulations with 0.5~\Mpch{} side boxes, labelled as Run05.0, Run05.100, Run05.1000, respectively.
}
\label{fig:Ngas}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{./Figure/Nstar/check_mfct_fnl_stars_compare_cumulative_lb027.ps}\\
\includegraphics[width=0.45\textwidth]{./Figure/Nstar/check_mfct_fnl_stars_compare_cumulative_lb039.ps}\\
\caption[Ngas]{\small
Cumulative cloud distributions as a function of stellar mass, for $z=8.51$ (upper panel) and $z=3.00$ (lower panel), for the runs with \fnl=0 (black solid lines), \fnl=100 (red long-dashed lines), and \fnl=1000 (blue dotted short-dashed lines).
We plot results for the high-resolution simulations with 100~\Mpch{} side boxes, labelled as Run100.0, Run100.100, Run100.1000, respectively.
}
\label{fig:Nstar}
\end{figure}
After discussing the general behaviours of the baryons in the Universe, it is interesting to go through their clumping properties and to investigate how the resulting ``visible'' objects are affected.
So, we identify gas clouds and dark-matter haloes by means of a FoF algorithm, as already described in Sect.~\ref{Sect:simulations}.
\\
To probe the effects of non-Gaussianities on the small primordial gas clouds, in Fig.~\ref{fig:Ngas}, we show the cumulative distributions at early times for the $0.5\,\rm\Mpch$ side boxes.
We focus on the \fnl=0, \fnl=100, and \fnl=1000 cases, at redshift $z=25$ and $z=15$.
At $z\sim 25$, the primordial non-Gaussian perturbations have well visible consequences in the gas clumpiness and lead to more massive and more numerous clouds with respect to the standard Gaussian scenario, for gaseous structures with mass $<10^{7}\msunh$ .
The differences are about a factor of $\sim 2-3$ in the \fnl=1000 case, but only $\sim 10$ per cent for the \fnl=100 case.
This reflects the larger probability at higher \fnl{} of forming objects that are able to trap and condense primordial cosmic gas.
The following growth of structures washes out the differences in a few $\sim 10^8\,\rm Myr$, and by $z\sim 15$ the gas distributions for these small objects are almost undistinguishable.
\\
The following evolution will lead to the formation of protostellar cores, to the birth of the first generation of stars, and then to the establishment of the standard stellar populations (see previous discussion).
To account for the differences which persist down to lower redshift on the stellar evolution properties, we plot in Fig.~\ref{fig:Nstar} the stellar mass distributions in the range $\sim 10^7-10^{13}\msun$, at redshift $z=8.51$ and $z=3$.
We consider the \fnl=0, \fnl=100, and \fnl=1000 cases in the large boxes, because they have higher-mass samples.
We find that also the stellar distributions are affected, mostly at earlier times, with differences that can reach a factor of $\sim 10$ for the \fnl=1000 case and a factor of a few for the \fnl=100 case, in the high-mass tail ($> 10^{9}\,\rm\msunh$).
At the low mass-end ($\sim 10^{7}-10^{8}\,\rm\msunh$), instead, differences are only a factor of $\sim 2$ for \fnl=1000, and $\sim 10$ per cent for \fnl=100.
At low redshift, $z\sim 3$, some differences persist for masses larger than $\sim 10^{11}\,\rm\msunh$, but the most of the objects shows well converging trends.
Below redshift $\sim 3$, the residual differences are further washed out by the subsequent structure formation and the on-going mass assembly process, so at $z\sim 0-1$ there are no distinctions at all in the various stellar components.
\section{Discussion and conclusions}\label{Sect:discussion}
\begin{table}
\centering
\caption[Simulation set-up]{Observational determinations of the \fnl{} parameter, according to different authors and different techniques.}
\begin{tabular}{lcc}
\hline
Reference & Range for \fnl & Confidence level\\
\hline
\cite{Komatsu2002} & $[-3500, +2000]$ & $2\sigma$\\
\cite{Komatsu2003} & $[-58, +134]$ & $2\sigma$\\
\cite{Gaztanaga2003}& $[+4, +18]$ & $1\sigma$\\
\cite{Gaztanaga2003}& $[-4, +5]$ & $1\sigma$\\
\cite{Chen2005} & $[-260, +40]$ & $1\sigma$\\
\cite{Chen2006} & $[-30, +74]$ & $1\sigma$\\
\cite{Cabella2006} & $[-80, +80]$ & $1\sigma$\\
\cite{Cabella2006} & $[-160, +160]$ & $2\sigma$\\
\cite{Spergel2007} & $[-54, +114]$ & $2\sigma$\\
\cite{Afshordi2008} & $[+109, +363]$ & $2\sigma$\\
\cite{Slosar2008} & $[-29, +70]$ & $2\sigma$\\
\cite{Slosar2008} & $[-65, +93]$ & $3\sigma$\\
\cite{Slosar2008} & $[-31, +70]$ & $2\sigma$\\
\cite{Slosar2008} & $[-96, +96]$ & $3\sigma$\\
\cite{Hikage2008} & $[-70, +91]$ & $3\sigma$\\
\cite{Yadav2008} & $[+27, +147]$ & $2\sigma$\\
\cite{Komatsu2009} & $[-9, +111]$ & $2\sigma$\\
\cite{Smith2009} & $[-4, +80]$ & $2\sigma$\\
\cite{Curto2009} & $[-8, +111]$ & $2\sigma$\\
\cite{Curto2009b} & $[-18, +80] $ & $2\sigma$\\
\cite{Vielva2009} & $[-94, +154] $ & $2\sigma$\\
\cite{Rudjord2009} & $[+44, +124]$ & $2\sigma$\\
\cite{Rudjord2010} & $[+42, +104]$ & $2\sigma$\\
\cite{Hou2010} & $[+34, +120]$ & $1\sigma$\\
\cite{Hou2010} & $[+12, +82]$ & $1\sigma$\\
\cite{Cabella2010} & $[-9, +85]$ & $1\sigma$\\
\cite{Komatsu2010} & $[+11, +53]$ & $1\sigma$\\
\cite{Komatsu2010} & $[-10, +74]$ & $2\sigma$\\
\hline
\label{tab:nG}
\end{tabular}
\begin{flushleft}
\vspace{-0.5cm}
{\small
}
\end{flushleft}
\end{table}
Cosmological structure formation is strongly affected by the initial perturbations which it originates from.
As a consequence of the central-limit theorem, it is common to assume a Gaussian distribution, but detailed studies of the CMB leave some room for deviations from the expected behaviour, and for an \fnl{} parameter different from zero.
Since non-Gaussianities alter the ``initial conditions'' of the problem of structure formation, and, thus, influence the temporal evolution of the forming structures, we have addressed their impacts on the visible, baryonic matter.
\\
Given the lack of extensive works on this subject, we have performed the first high-resolution cosmological N-body, hydrodynamical, chemistry simulation of cosmic structure formation, in the frame of the $\Lambda$CDM model and in presence of primordial non-Gaussianities.
We have studied the principal properties of gas and stars in different scenarios, by following: cosmic evolution of e$^-$, H, H$^+$, H$^-$, He, He$^+$, He$^{++}$, H$_2$, H$_2^+$, D, D$^+$, HD, HeH$^+$ \cite[][]{Yoshida2003,Maio2006,Maio2007,Maio2009,Maio2010a}; radiative gas cooling at high \cite[][]{SD1993} and low temperatures \cite[both from molecules and fine-structure metal transitions, according to ][]{Maio2007}; multiphase star formation \cite[][]{Springel2003}; UV background radiation \cite[][]{HaardtMadau1996}; wind feedback \cite[][]{Springel2003,Aguirre_et_al_2001}; metal pollution from popIII and/or popII stellar generations, ruled by a critical metallicity threshold of $Z_{crit}=10^{-4}\,\zsun$ \cite[][]{Tornatore2007,Maio2010a} and leading gravitational enrichment into the surrounding medium \cite[][]{Maio2011arXiv}.
\\
Observational evidences and continuous investigations over the last decade (see details in Tab.~\ref{tab:nG}) have progressively constrained \fnl{} to positive values, thus, we have considered \fnl=0, 10, 50, 100, and 1000.
From Tab.~\ref{tab:nG}, the most likely values\footnote{
In Tab.~\ref{tab:nG}, the CMB notation is assumed, thus, to compare with the values adopted in the simulations (\fnl=0, 10, 50, 100, 1000), one has to multiply the tabulated data by $\sim 1.28$.
}
are expected to be between \fnl$\sim 0-100$, and the \fnl=1000 case has to be simply considered as an extreme upper limit.
\\
The goal of this work has been to theoretically explore the effects of different \fnl{} values on the structure formation properties and on the baryon history.
Thanks to the high resolution reached (up to $\sim 40\,\msunh$), we are able to address the cosmic evolution from the very early times -- when the first star formation episodes take place -- till to the present.
\\
Our main results (Sect.~\ref{Sect:results}) have shown that structure formation processes (Sect.\ref{Sect:maps}), star formation history (Sect.~\ref{Sect:SFR}), chemical evolution (Sect.~\ref{Sect:chemistry}), and visible-matter distributions (Sect.~\ref{Sect:distributions}) are all affected by non-Gaussianities.
However, while the effects are quite evident for the extreme \fnl=1000 case, they are much milder for smaller, more ralistic values of \fnl.
\\
Indeed, gas temperature and composition reveal some effects of primordial non-Gaussianities only at very high redshift, with changes of the order of $\sim 10$ per cent, at $z > 10$, for \fnl=$0-100$ and of a factor of $\sim 2$ for \fnl=1000 (Fig.~\ref{fig:Maps}).
\\
Primordial molecular fractions (Fig.~\ref{fig:Maps2} and Fig.~ \ref{fig:Maps3}) are particularly senstive to the thermodynamical properties of the medium and their behaviour increases mostly in the high-density regions, where the effects on clustering due to primordial non-Gaussianities are more relevant.
\\
Star formation history (Fig.~\ref{fig:SFR}) has a converging behaviour despite of non-Gaussianities. The main discrepancies among the models are found for the onset, which takes place at earlier epochs for larger \fnl.
The time-lags are around some $10^7\,\rm yr$, at $z\sim 15$, and they are reflected in the transition from the popIII regime, described by a top-heavy IMF, to the popII regime, described by a Salpeter IMF (Fig.~\ref{fig:SFR}).
Despite such discrepancies, the popIII contributions stay around $\sim 10^{-3}-10^{-2}$ at $z\sim 10$ for all the models, with only $\sim 1-10$ per cent scatter.
\\
Given the uncertainties on the primordial popIII IMF we also run corresponding simulations with Salpeter-like popIII IMF and found that (Fig.~\ref{fig:SFR_IMFrange}) the resulting deviations of the popIII star formation rates could be larger than the non-Gaussian ones.
\\
Additionally, we compared with the second-order corrections in linear perturbation theory \cite[][]{TseliakhovichHirata2010} and their impact on structure formation evolution \cite[][]{Maio2011}, showing that (Fig.~\ref{fig:SFR_vb}) also the resulting primordial gas bulk flows could have comparable or larger repercursions on the visible matter.
Therefore, the behaviour of baryons seems tightly bound and twisted with several physical mechanisms, whose effects are degenerate with non-Gaussianities.
\\
As a consequence of the trends of the star formation rates, metal enrichment results (Fig.~\ref{fig:Zcompare}) are affected by the initial \fnl, mostly at early times, with pollution histories converging at low redshift ($z\lesssim 12$) and overcoming, in average, the critical metallicity at $z\sim 6$, when the Universe has an age of roughly 1~Gyr.
This is quite independent from the primordial non-Gaussianities, popIII IMF, or additional second-order effects, but at high redshift metal spreading is highly sensitive to all these phenomena and can show degeneracies with them.
\\
Gaseous and stellar mass statistics (Fig.~\ref{fig:Ngas}, and Fig.~\ref{fig:Nstar}) show variations at the high-mass tail, and rare, more massive objects are more common in higher-\fnl{} cosmologies.
The distributions of the very first structures show an increment of a factor of a few in the \fnl=1000 case, but only some $\sim 10$ per cent for \fnl$=0-100$, mainly at the high-mass end.
However, the subsequent evolution at later times washes them out almost completely.
\\
Throughout this work, we have used the \fnl{} parameter only to quantify deviations from the Gaussian shape.
We have not investigated models with a negative \fnl{} as they have been progressively dismissed by observations.
However, for sake of completness, we note that, theoretically, one would expect roughly specular behaviours with respect to the \fnl=0 case when universes with negative \fnl{} are considered -- a result already well established from simple dark-matter-only simulations (as mentioned in Sect.~\ref{Sect:introduction}).
Indeed, in such scenarios, the probability distribution function of the initial density field would be skewned towards the underdense tail, causing a paucity of the more massive objects and, in general, a delayed evolution of the cosmic structures.
We therefore expect the variations from the \fnl=0 case originated from a positive \fnl{} parameter to be accordingly reversed when the corresponding negative \fnl{} parameter is adopted.
\\
Besides the \fnl{} formulation, higher-order parameters can be introduced, but their uncertainties are huge with, e.g., $-7.4\times 10^5<$\gnl$<8.2\times 10^5$ \cite[][]{Smidt2010}.
Thus, a study of their impacts on the cosmic structures would be quite vague and elusive.
Furthermore, the effects of the second-order \fnl{} parameter are quite small, so it is reasonable to think that inclusion of higher-order parameters will not dramaticaly change our conclusions.
Even if future initial conditions will include additional higher-order corrections, being definitely more complete and detailed, probably they would not alter the global scenario we have drawn.
For what we said in the previous sections, other physical phenomena, like
cosmic rays\footnote{
See e.g. \cite{Ginzburg1964,Mannheim1994,Biermann2001,Sigl2004,Stanev2004,Vasiliev2006,Pfrommer2007,Jasche2007}.
},
radiative transfer and reionization\footnote{
See e.g. \cite{Abel1999,Gnedin2001,Bene2001,Bene2003,Maselli2003,Whitehouse2005,Iliev2006,Mellema2006,Trac2007,Aubert2008,Johnson2008,Pawlik2008,Finlator2009,Margarita2009,Margarita2010,Cantalupo2011,Wise2010arXiv}.
},
magnetic fields\footnote{
See e.g. \cite{Orszag1979,Phillips1985,Brio1988,Dolag1999,Balsara1999,Londrillo2000,Toth2000,Borve2001,Brueggen2005,Price2005,Li2008,Stone2008,Ryu2008,Dolag2009,Brandenburg2010,Burzle2010arXiv,Gaburov2010arXiv,Kotarba2010}.
},
AGN feedback\footnote{
See e.g. \cite{Begelman1978,Blandford1982,Rees1984,Ciotti1997,King2003,Sazonov2005,Churazov2005,Granato2006,Ciotti2007,Khochfar2008,DiMatteo2008,Sijacki2009,Milo2009,McKernan2010,Volonteri2010arXiv,Teyssier2010arXiv,McNamara2011}.
},
etc., are not supposed to experience strong modifications, because of primordial non-Gaussianities.
\\
We stress that even small variations in the determinations of the cosmological parameters, above all errors of a few per cent in the estimates of $\sigma_8$, could add further degeneracies to disentangle the purely non-Gaussian results from the primordial IMF, gas bulk flows, etc.
\\
We summarize our conclusions by stating that for reasonable values of \fnl$=0-100$ in the primordial non-Gaussianities (in agreement with observations):
\begin{itemize}
\item
early molecular fractions in the cold, dense gas phase are predicted to be altered of $\sim 10$ per cent;
\item
small temperature fluctuations of $\,\lesssim 10$ per cent are induced during the first Gyr of the cosmic evolution;
\item
the onset of the first star formation events, at $z\gtrsim 15$, is influenced of some $10^7\,\rm yr$;
\item
the epoch of popIII/popII transition is affected by some $10^7\,\rm yr$;
\item
variations of $\lesssim 10$ per cent are found in the mass distributions during the formation of first structures;
\item
only mild variations in the chemical history of the Universe are expected;
\item
extremly high values for \fnl{} would imply corresponding variations of a factor $\sim 2-3$ in the thermodynamical and chemical properties of the gas, earlier star formation histories of $\sim 0.5-1\times 10^8\,\rm yr$, and changes up to a factor of $\sim 10$ in the high-mass end of the gas cloud and stellar mass distributions at early times;
\item
non-Gaussian effects could have degeneracies with other physical processes, like primordial gas bulk flows, with unknown star formation quantities, like popIII IMF, or with the determination of cosmological parameters, mostly $\sigma_8$ at a few per cent level.
\end{itemize}
\section*{acknowledgments}
We acknowledge useful discussions with K.~Dolag, M.~Grossi, and L.~Moscardini.
UM also acknowledges B.~Ciardi, S.~Khochfar, and G.~Rossmanith.
We also thank the referee M.~Stiavelli for his useful and prompt comments on the manuscript.
The simulations were performed on the IBM Power~6 machines of the Garching Computing Center (Rechenzentrum Garching) of the Max Planck Society (Max-Planck-Gesellschaft).
\bibliographystyle{mn2e}
|
\section{Introduction}
\label{sec:intro}
A Constraint Satisfaction Problem (CSP) consists of $N$ variables,
$ {\vec{x}} :=(x_1,\ldots,x_N)$, and $M$ clauses, i.e. $\{0,1\}$-valued
functions, $(\Phi_1( {\vec{x}} ),\ldots,\Phi_M( {\vec{x}} ))$. An assignment $ {\vec{x}} $
is a solution if all clauses simultaneously evaluate to $1$. In the
context of wireless networks, we show that CSPs provide a unifying
framework that encompasses many important resource allocation tasks.
Examples include: allocation of radio channels so that transmissions
by neighbouring WLANs (Wireless Local Area Networks) or mobile
phone cells do not interfere; the selection of packets to be XORed
on each link in network coding; and finding a non-colliding schedule
of time-slots in a WLAN. Unlike in traditional CSPs, however, in
network applications each constrained variable $x_i$ is typically
co-located with a physically distinct device such as an access
point, base-station or a link. The resulting communication constraints
impose severe restrictions on the nature of the algorithm that can
be used for solving the CSP. These restrictions are violated by
existing CSP solvers, leading us to define a new class of algorithms
that we term \emph{decentralized CSP solvers}. We constructively
establish the existence of decentralized solvers by introducing a
family of randomized algorithms belonging to this class.
Roughly speaking, decentralized CSP solvers must be capable of
finding a satisfying assignment, $ {\vec{x}} $, while updating each variable
$x_i$ based solely on knowledge of whether all of the clauses in
which $x_i$ participates are satisfied or at least one clause is
unsatisfied. The following example gives a concrete illustration
of the origins of these constraints.
\subsection{Motivating example: channel allocation}
\label{sec:channel_all}
802.11 WLANs are ubiquitous, but their deployments are unstructured
and they operate in an unlicensed frequency band. Within this band,
an 802.11 WLAN can select one of several channels, typically $11$,
to operate on. In the vast majority of WiFi routers channel
selection is, at present, based on operator selection without any
quantitative instruction from the router. Self-selection of this
channel is the task we consider here. Co-ordinated selection is
hampered by the fact that the interference range of a typical 802.11
device is substantially larger than its communication range.
Consequently, WLANs can interfere but may be unable to decode each
others messages (illustrated schematically in Fig. \ref{fig:example}).
Discovery by a WLAN of the existence of interfering WLANs via its
wired back-haul may be prevented by firewalls and, in any case, these
WLANs may not know their physical location. These practical
restrictions mandate a decentralized channel-selection algorithm.
We can identify the task of each WLAN self-selecting a non-interfering
channel with a CSP with communication constraints imposed by the
network topology. Let $x_i$ be the channel selected by WLAN
$i\in\{1,\ldots,N\}$ and define $M=N(N-1)/2$ clauses, one for each
pair of WLANs, that evaluates to one if the WLANs are on non-interfering
channels or are out of interference range and is zero otherwise.
Due to the unstructured nature of the deployment, the lack of shared
administrative control and communication constraints, a WLAN cannot
rely on knowing the number or the identity of interfering WLANs or
the channels that they currently have selected. A practical CSP
solver for this task can only rely on each WLAN being able to measure
whether: (i) all of the neighbouring WLANs have selected a different
channel from it; or (ii) if one or more neighbours have selected
the same channel.
\begin{figure}
\centering
\includegraphics[width=0.8\columnwidth]{example}
\caption{An illustration of overlapping interference regions in
neighbouring WLANs. Black dot indicates location of a wireless
access point, shaded area indicates region within which communication
with access point can take place and outer circles indicate regions
within which transmissions interfere.}
\label{fig:example}
\end{figure}
In certain settings, a limited amount of communication of variable
values between the stations hosting them may be feasible. For
example, it may be possible for stations hosting a link or for a
WLAN to overhear beacons or traffic from a subset of its interferers,
or for an access point to communicate with other, nearby access
points. Algorithms for channel selection that are proposed in the
literature assume the existence of end-to-end communication for
centralized solutions, for message passing in decentralized or
gossiping solutions, or to co-ordinate global restart in simulated
annealing proposals e.g. \cite{Wu_06, Kauffmann07, Crichigno08,
Subramanian08} and references therein. This information, however,
tends to be opportunistic in nature and cannot be relied on \emph{a
priori}. We leave how best to exploit such partial information to
future work, focusing here on the most challenging, fundamental
cases where fully decentralized algorithms are required.
\subsection{Contribution}
The primary contributions of this paper are fourfold. Firstly, we show
that CSPs provide a unifying framework that encompasses several
important problems in wireless networks, including channel selection,
inter-session network coding and decentralized scheduling in
a WLAN. Each of these problems is generally thought of as being
distinct and requiring different solution approaches. We show
that they are fundamentally related as CSPs so that the decentralized
CSP solver introduced in this paper can be used for all of them.
Secondly, we define a new class of CSP algorithms which we term
decentralized CSP solvers and, extending our earlier approach to
graph colouring \cite{leith06}\cite{duffy08}, introduce a
novel stochastic decentralized CSP solver proving that it will find
a solution in almost surely finite time, should one exist, and
showing it has many practically desirable properties. Thirdly, we
benchmark the algorithm's performance on a well-studied class of
application-agnostic CSPs, random k-SAT. For problems with up to
a thousand variables, which are large for our motivating examples,
we find that the time the algorithm takes to find a satisfying
assignment is close to that of WalkSAT \cite{Selman95}, a well
regarded, efficient centralized CSP solver, while also possessing
desirable features of Survey Propagation \cite{Braunstein05}. That
is, despite its decentralized nature, the algorithm is fast. Fourthly,
we demonstrate the solver's practical utility in a complex, wireless
resource allocation case study.
The rest of the paper is organized as follows. In Section
\ref{subsec:dCSPs} we define decentralized CSP solvers and show
that well-known algorithms for solving CSPs fail to meet these
criteria. In Section \ref{sec:CFL} we introduce a decentralized
CSP solver. We prove that it finds a solution in finite time with
probability one whenever a feasible solution exists, obtaining an
upper bound on the algorithm's convergence rate in the process. In
Section \ref{sec:sim} we consider its speed at identifying solutions
of instances of random k-SAT. In Section \ref{sec:example} we
demonstrate the algorithm's utility in solving the problems that
motivated its introduction. Section \ref{sec:disc} contains discussion
and closing remarks.
\section{Decentralized CSP solvers}
\label{subsec:dCSPs}
We begin by formalizing the criteria that CSP solvers must possess
to be of use for problems with communication constraints such as
those outlined above. We call algorithms that meet these criteria
decentralized CSP solvers.
\begin{definition}[CSP]
A CSP with $N$ variables, $\{x_1,\ldots,x_N\}$, and $M$ clauses is
defined as follows. The variables each take values in a finite set
$ {\mathcal{D} }=\{1,\ldots,D\}$
and $ {\vec{x}} :=(x_1,\ldots,x_N)\in {\mathcal{D} }^N$. Each clause $m\in {\mathcal{M} }=\{1,\ldots,M\}$
is defined by a function $\Phi_m: {\mathcal{D} }^N\mapsto\{0,1\}$ where for an
assignment of variables, $ {\vec{x}} $, $\Phi_m( {\vec{x}} )=1$ if clause $m$ is
satisfied and $\Phi_m( {\vec{x}} )=0$ if clause $m$ is not satisfied. An
assignment $ {\vec{x}} $ is a solution to the CSP if all clauses are
simultaneously satisfied. That is,
\begin{align}
\label{eq:CSP}
{\vec{x}} \text{ is a satisfying assignment iff }
\min_{m\in {\mathcal{M} }} \Phi_m( {\vec{x}} )=1.
\end{align}
\end{definition}
This encompasses all of our examples of interest as well as k-SAT,
the most well-studied general class of CSPs.
A CSP solver is an algorithm that can find a satisfying variable
assignment for any solvable CSP.
\begin{definition}[CSP Solver] A CSP solver takes a CSP as input
and determines a sequence $\{ {\vec{x}} (t)\}$ such that for any CSP that
has satisfying assignments:
\begin{description}
\item[(D1) ]
for all $t$ sufficiently large $ {\vec{x}} (t)= {\vec{x}} $ for some satisfying
assignment $ {\vec{x}} $;
\item[(D2) ]
if $t'$ is the first entry in the sequence $\{ {\vec{x}} (t)\}$ such that
$ {\vec{x}} (t')$ is a satisfying assignment, then $ {\vec{x}} (t)= {\vec{x}} (t')$ for
all $t>t'$.
\end{description}
\end{definition}
To provide criteria that classify CSP solvers as being decentralized,
we begin with the following definition.
\begin{definition}[Clause participation]
We say that variable $x_i$ participates in clause $\Phi_m( {\vec{x}} )$ if
the value of $x_i$ can influence the clause's satisfaction for at
least one assignment of the rest of the variables in $ {\vec{x}} $.
\end{definition}
For each variable $x_i$, let $ {\mathcal{M} }_i$ denote the set of clauses in
which it participates:
\begin{align*}
{\mathcal{M} }_i =
\bigcup_{x_i\in\{1,\ldots, d\}} \left(
\bigcup_{j\neq i} \bigcup_{x_j \in \{1,\ldots, d\}}
\left\{m: \Phi_m(x_i, \{x_j\}) = 0 \right\}\right. \\
\bigcap
\,
\left.
\bigcup_{j\neq i} \bigcup_{x_j \in \{1,\ldots, d\}}
\left\{m: \Phi_m(x_i, \{x_j\}) = 1 \right\}
\right).
\end{align*}
\begin{comment}
\begin{align*}
{\mathcal{M} }_i = \bigcup_{j\neq i} \bigcup_{x_j \in \{1,.., d\}}
&\left\{m:
\min_{x_i \in \{1,\ldots, d\}} \Phi_m(x_i, \{x_j\}) = 0 \right. \\
&\left.
\qquad \text{ and }
\max_{x_i \in \{1,\ldots, d\}} \Phi_m(x_i, \{x_j\}) = 1
\right\}.
\end{align*}
\end{comment}
Re-writing the left hand side of eq. \eqref{eq:CSP}
in a way that focuses on the satisfaction of each variable we have:
\begin{align}
\label{eq:dCSP}
{\vec{x}} \text{ is a satisfying assignment iff } \min_i \min_{m\in {\mathcal{M} }_i}
\Phi_m( {\vec{x}} ) = 1.
\end{align}
A decentralized CSP algorithm can be thought of as having
intelligence co-located with each variable. The intelligence at
variable $x_i$ can only determine whether all of the clauses that
$x_i$ participates in are satisfied or that at least one clause is
unsatisfied and must make update decisions locally based solely on this
knowledge.
\begin{definition}[Decentralized CSP Solver]
A decentralized CSP solver is a CSP solver that for each variable
$x_i$, $i\in\{1,\ldots,N\}$, must select its next value based only
on an evaluation of
\begin{align}
\label{eq:minPhi}
\min_{m\in {\mathcal{M} }_i} \Phi_m( {\vec{x}} ).
\end{align}
That is, for each variable $x_i$ the solver must make a decision
on the next value of $x_i$ based solely on knowing whether that
variable is currently satisfied or not without explicitly knowing:
\begin{description}
\item[(D3) ]
the assignments of other variables, $x_j$ for any $j\neq i$;
\item[(D4) ]
the set of clauses that any variable, including itself, participates
in, $ {\mathcal{M} }_j$ for $j\in\{1,\ldots,N\}$;
\item[(D5) ]
the functions that define clauses, $\Phi_m$ for $m\in\{1,\ldots,M\}$.
\end{description}
\end{definition}
Note that the properties (D1) (D2) of any CSP solver
mean that a decentralized CSP solver must, without explicit
communication, settle upon a satisfying assignment the first time
one is found. Communication to co-ordinate global stopping or
restarting of the solver would be contrary to the nature of the
natural constraints of these problems and so is forbidden.
The stochastic algorithm we define in Section \ref{sec:CFL} shall
provably satisfy the properties of a decentralized CSP solver with
probability one. Moreover, it will also have the desirable property
that it will automatically restart its search, without communication,
upon any change to the clauses of a CSP that makes the current
variable assignment no longer satisfying. This is significant in
practical applications as, for example, the arrival of new transmitters
in a wireless network would induce a change the associated CSP.
\subsection{Formulating Wireless Network Tasks As CSPs}
Before proceeding, we demonstrate that several important resource
allocation tasks in wireless networks fall within this CSP framework,
which, therefore, provides a unifying framework for analyzing these
tasks.
{\it 1) Graph colouring.}
As briefly described in Section \ref{sec:channel_all}, the channel
assignment problem corresponds to decentralized graph colouring.
In the simplest model, the network is described by an undirected
interference graph $G=(V,E)$ with vertices $V$ and edges $E$. Each
vertex represents a WLAN and an edge exists between two WLANs if
they interfere with each other when on the same channel.
Define $N=|V|$ and $M=|E|$ and let $ {\mathcal{D} }$ denote
the set of available colours (channels) and let $x_i$ be a variable
with value equal to the colour selected by an intelligence co-located
with each vertex $i\in V$. Each clause $m\in\{1,\ldots,|E|\}$, which
is an enumeration of the elements of $E$, can be identified with
an edge $(i,j)\in E$. We define
\begin{align*}
\Phi_m( {\vec{x}} )= \Phi_m(x_i,x_j)=
\begin{cases}
1 & \text{if } x_i\ne x_j\\
0 & \text{otherwise}.
\end{cases}
\end{align*}
The participation set, $ {\mathcal{M} }_i$, of a variable $x_i$ consists of
all clauses where the vertex $i$ is one end of the associated edge,
\begin{align*}
{\mathcal{M} }_i = \left\{m\equiv (i,j): (i,j)\in E\right\}.
\end{align*}
A variable assignment satisfies this CSP if and only if it is
also a proper colouring for the graph $G$. Proper
colourings correspond to channel assignments in which no two
neighbouring WLANs in the interference graph have selected the same
channel. To find a proper colouring, when one exists, a decentralized CSP
solver requires only that an intelligence co-located with each
vertex $i$ can measure whether: (i) all of its neighbours are using
a different colour from vertex $i$; or (ii) at least one neighbour
has selected the same colour as vertex $i$. This information is
sufficient to evaluate eq. \eqref{eq:minPhi} and, in particular,
the intelligence at each vertex does not need to know: (D3) the
colour selected by any other vertex; (D4) its set of neighbours;
or (D5) the exact nature of the clauses $\Phi_m$, $m\in\{1,\ldots,M\}$.
{\it 2) Channel assignment with channel-dependent interference.}
Whether or not transmitters interfere in the channel assignment task
may depend on the radio channel selected. This can arise due to
frequency dependent radio propagation or to channel dependent
spectral masks. That is, for regulatory reasons, different spectral
masks are typically used when transmitting on channels at the edge
and on channels in the middle of a radio band. This problem then
has a collection of conflict graphs, one for each possible radio
channel. This version of the channel assignment task can also be
formulated as a CSP even though it is no longer a graph colouring
problem. Let $G(c)=(V,E(c))$, $c\in {\mathcal{D} }$ be a set of undirected
graphs with the same vertices $V$ but possibly differing edge sets
$E(c)$. Again each vertex represents a WLAN, but the interference
graph is channel-dependent.
The Graph $G(c)$ is associated with radio channel $c$ and
an edge exists in $E(c)$ if two WLANs interfere while on channel
$c$. Let $x_i$ be a variable with value equal to the channel selected
by each vertex $i\in V$. Each clause
$m\in\{1,\ldots,|E(1)|+\cdots+|E(D)|\}$, which is an enumeration of
all edges in all graphs, can be identified with a $c\in {\mathcal{D} }$ and an
edge $(i,j)\in E(c)$ and is defined by
\begin{align*}
\Phi_m( {\vec{x}} )= \Phi_m(x_i, x_j)=
\begin{cases}
0 & \text{if } x_i = x_j = c\\
1 & \text{ otherwise.}
\end{cases}
\end{align*}
The collection of clauses that variable $x_i$, associated with
vertex $i$, participates in can be identified as
\begin{align*}
{\mathcal{M} }_i = \bigcup_{c\in {\mathcal{D} }}\left\{m\equiv (i,j,c): (i,j)\in E(c)\right\}.
\end{align*}
With $x_i$ being the current channel selection of vertex $i$, for
all $c\neq x_i$ the clauses associated with $m\equiv (i,j,c)$ are
automatically satisfied. Thus to evaluate eq. \eqref{eq:minPhi}
it is sufficient for the station to measure if no neighbour coincides
with its current channel selection or if at least one does.
To find an interference-free channel assignment, a decentralized
CSP solver requires that each vertex $i$ can measure whether: (i)
all of its neighbours on the currently selected channel are using
a different channel from itself; or (ii) at least one neighbour has
selected the same channel.
{\it 3) Inter-session network coding.}
Network coding has been the subject of considerable interest in
recent years as it offers the potential for significant increases
in network capacity \cite{Koetter03, Deb05, Katti06}. In network
coding, network elements combine packets together before transmission
rather than forwarding them unmodified. The combined packets can
be from individual flows, known as intra-flow coding, or across
multiple flows, inter-flow coding.
While intra-flow coding within multicast flows has been well studied,
inter-flow coding between unicast flows has received less attention,
yet is, perhaps, of more immediate relevance to current Internet
traffic. Inter-flow coding is known to be challenging \cite{Sengupta10}.
The task of a network finding a feasible linear network code
in a distributed fashion, but with some global sharing of calculations,
is investigated in \cite{Kim09} through the use of a genetic
algorithm. This task can be formulated as a CSP.
Let $G=(V,E)$ denote a directed acyclic multi-graph representing
the network with vertices $V$ and edges $E$. Each vertex
represents either a source of an information flow, a destination
or router in the network. Edges represent physical connections
between these elements. Time is slotted and each edge can
transmit a single packet per slot. We allow multiple edges
between vertex pairs in order to accommodate higher rate links. It
is the goal of each link beyond the source to determine a linear
combination of its incoming packets to forward so that ultimately
all flows get to their destinations. As the combination of packets
coded by each link will take care of routing, each flow $p\in P$
is defined by its source, $\sigma(p)\in V$, and its destination
$\delta(p)\in V$. Each flow is assumed to be unit rate with higher
rate flows accommodated by flows with the same source and destination
vertices. We will treat the collection of source, $\bigcup_{p\in
P}\{\sigma(p)\}$, and destination vertices, $\bigcup_{p\in
P}\{\delta(p)\}$, as special, solely having one incoming and one
outgoing edge per flow respectively.
For each edge $i\in E$, define $s(i)$, $s:E\mapsto V$, to be its
source vertex and $t(i)$, $t:E\mapsto V$, to be its target vertex.
For each vertex $v\in V$ we define the set of incoming edges
$I_v:=\{i\in E: t(i)=v\}$ and the set of outgoing edges $O_v:=\{i\in
E: s(i)=v\}$. For each edge not corresponding to source, $i\in
E\setminus \bigcup_{p\in P}O_{\sigma(p)}$, we associate a variable
$x_i\in {\mathcal{D} }:=\{1,\ldots,2^{|P|}\}$, which is an enumeration of
the power set of the set of flows $P$. We define a bijective map
$\psi: {\mathcal{D} }\mapsto\{0,1\}^{|P|}$ that determines a vector whose
positive entries correspond to flows to be coded.
In total, we have $|E\setminus\bigcup_{p} O_{\sigma(p)}|$ clauses.
One for every link that isn't an outgoing link from a source vertex.
At each time slot, each link that doesn't correspond to a source or
final destination link,
$i\in E\setminus\bigcup_{p}(O_{\sigma(p)}\cup I_{\delta(p)})$,
wishes to forward a packet consisting of XORed packets (corresponding
to addition in the Galois field of two elements) from the flows
indicated by $\psi(x_i)$. Each final link $i\in\bigcup_{p}
I_{\delta(p)}$ wishes to forward packets from $p$, which we
indicate by the vector $\gamma_p$ with a $1$ at the location $p$
and zeros elsewhere. A link and its immediate upstream neighbouring
links will be dissatisfied if it cannot do so.
More formally, for each edge not corresponding to
the outgoing link from a source or incoming link to a destination,
$i\in E\setminus \bigcup_{p\in P}(O_{\sigma(p)}\cup I_{\delta(p)})$,
we have a clause $m\equiv i$, $\Phi_m( {\vec{x}} )$, defined by
\begin{align*}
\Phi_m(x_i,x_{j:j\in I_{s(i)}})
=
\begin{cases}
1 & \text{if } \exists\, \theta \text{ s.t. }
\Psi(I_{s(i)}) \theta^T= \psi(x_i)^T\\
0 & \text{otherwise}.
\end{cases}
\end{align*}
where $\theta$ is a binary vector and $\Psi(I_{s(i)})$ is the
rectangular matrix consisting of $\psi(x_j)^T$ for $j\in
I_{s(i)}\setminus\bigcup_{p\in P}O_{\sigma(p)}$ and
and $\gamma_p^T$ if $i\in O_{\sigma(p)}$ for some $p$.
To complete the CSP, we need to ensure that packets from flow
$p$ can be decoded at their destination $\delta(p)$. For
$i\in\bigcup_{p\in P}I_{\delta(p)}$, then with $\delta(p')=t(i)$
we introduce a clause
\begin{align*}
\Phi_m(x_{j:j\in I_{s(i)}})
=
\begin{cases}
1 & \text{if } \exists\, \theta \text{ s.t. }
\Psi(I_{s(i)})\theta^T = \gamma_{p'}^T\\
0 & \text{otherwise}.
\end{cases}
\end{align*}
where again $\theta$ is a binary vector and $\Psi$ is is the
rectangular matrix defined above. The set of clauses a link variable
$x_i$, $i\in
E\setminus (\bigcup_{p\in P}O_{\sigma(p)} \cup I_{\delta(p)})$,
participates in is
\begin{align*}
{\mathcal{M} }_i =\{ m \equiv j: j=i \text{ or } i\in I_{s(j)}\}
\end{align*}
Any variable assignment then satisfies this CSP if and only if it
is a realizable network code satisfying all flow demands.
Hence, to find a proper assignment a decentralized CSP
solver requires only that each link $i$ can determine whether: (i)
its own coding is realizable and the coding of its immediately
down-stream links are realizable; or (ii) if at least one of these
is not realizable. This is sufficient to evaluate eq. \eqref{eq:minPhi}.
Each link does not need to explicitly know: (D3) the code selected
by any other link; (D4) the network topology; (D5) any details
of how code realizability is determined at any link.
{\it 4) Decentralized transmission scheduling.} In an Ethernet or
WLAN it is necessary to schedule transmissions by stations. This
might be achieved in a centralized manner using TDMA, but it can
also be formulated as a decentralized problem. The classical CSMA/CA
approach to decentralized scheduling never settles to a single
schedule and some comes at a cost of the possibility of continual
collisions. Recently, there has been interest in decentralized
approaches for finding collision-free schedules, see
\cite{barcelo10,fang12}. This task can be formulated as a CSP as
follows. Let $V$ denote the set of transmitters in the WLAN, $T$
denote the set of available time slots and define $N=|V|$ and
$M=N(N-1)/2$. Let $x_i$ be a variable with value equal to the
transmission slot selected by transmitter $i\in V$. Define a clause
$\Phi_m( {\vec{x}} )$ associated with each pair of transmitters $m\equiv
(i,j)$ such that
\begin{align*}
\Phi_m(x_i,x_j)=
\begin{cases}
1 & \text{ if } x_i\ne x_j \\
0 & \text{ otherwise}.
\end{cases}
\end{align*}
The participation sets are
\begin{align*}
{\mathcal{M} }_i=\bigcup_{j\neq i}\{m \equiv (i,j)\}.
\end{align*}
Any variable assignment satisfies
this CSP if and only if it is also a collision-free time-slot
schedule. To find a collision-free schedule, when one exists, a
decentralized CSP solver requires only that each transmitter $i$
can measure whether: (i) all of its neighbours are using a different
time-slot from transmitter $i$; or (ii) at least one transmitter
in the WLAN has selected the same time-slot as transmitter $i$.
Again, this is all that is needed to evaluate eq. \eqref{eq:minPhi}.
Each transmitter does not need to know: (D3) the time-slot selected
by any other transmitter; (D4) the set of transmitters; (D5) the
clauses.
\subsection{Related work - existing algorithms are not decentralized}
The literature on general purpose CSP solvers is vast, typically
focusing on solving k-SAT problems in conjunctive normal form, but
they can be broadly classified into those based on: (i) the
Davis-Putnam-Logemann-Loveland (DPLL) algorithm \cite{Davis60,
Davis62}; (ii) Survey Propagation \cite{Braunstein05}; and (iii)
on Stochastic Local Search (SLS) \cite{Selman95}. Each of these
approaches has experienced substantial development and has its own
merits, but none were motivated by problems where variables
have a geographical sense of locality.
Algorithms developed from the DPLL approach have proved to be the
quickest at SAT-Race and SAT Competition in recent years, e.g.
ManySAT \cite{Hamadi09}. The DPLL approach ultimately guarantees a
complete search of the solution space and so meets the (D1)
and (D2) criteria. They are, however, based on a branching
rule methodology, e.g. \cite{Ouyang98}, that assumes the existence
of a centralized intelligence that employs a backtracking search.
The implicit assumptions of the information available to this
intelligence breaks the conditions (D3) (D4) (D5) and so they are
not decentralized CSP solvers.
Survey propagation, a development of belief propagation \cite{pearl82}
from trees to general graphs, has proved effective in graphs that
do not contain small loops \cite{mezard02a}. For a given CSP, the
fundamental structure of study is a called a factor graph. In order
to generate this, it is necessary to know what clauses each variable
participates in and the nature of each of the clauses, breaking the
(D4) and (D5) criteria and so these are not decentralized CSP solvers.
SLS algorithms also depend fundamentally upon the exchange of
information, mostly in an explicit manner breaking the (D3) condition
by basing update decisions on relative rankings of the constraint
variables but also in a more subtle fashion. To see this implicit
requirement, consider the following algorithm for binary valued
variables originally proposed by Papadimitriou \cite{papadimitriou91}
and developed further by Sch\"oning \cite{schoning99}. Pick a random
assignment of values for the constraint variables. Repeat the
following: from all of the unsatisfied clauses, pick one uniformly at
random, select one of the variables participating in that clause
and negate its value, breaking the (D4) and (D5) conditions.
The algorithm halts when all clauses are satisfied or a specified time
limit expires. Although simple, this forms the basic building block
for all SLS algorithms, including the well-studied WalkSAT algorithm
\cite{Selman95}. It is important that a single unsatisfied clause
is selected at each step and that a single variable within the
clause is adjusted as it is this that leads to the algorithm behaving
as a random walk \cite{schoning99}. Thus, again, solvers in this
class are not decentralized.
\section{A decentralized CSP solver}
\label{sec:CFL}
We now introduce an algorithm that satisfies the decentralized CSP
solver criteria.
\subsection{Communication-Free Learning Algorithm}\
An instance of the following Communication-Free Learning (CFL)
algorithm is run in parallel for every variable. For each variable,
$i\in\{1,2,...,N\}$ it keeps a probability distribution,
$p_{i,j}$ over $j\in D$ as well as the current variable value $x_i$.
In pseudocode, the CFL algorithm is:
\begin{algorithm}
\caption{Communication-Free Learning}
\label{CFL}
\begin{algorithmic}[1]
\STATE Initialize $p_{i,j}=1/d$, $j\in\{1,...,D\}$.
\LOOP
\STATE
Realize a random variable, selecting $x_i=j$ with probability $p_{i,j}$.
\STATE
Evaluate $\min_{m\in {\mathcal{M} }_i} \Phi_m( {\vec{x}} )$, returning \emph{satisfied}
if its value is $1$, as this indicates all of variable $i$'s clauses
are satisfied given the present assignment, and \emph{unsatisfied}
otherwise.
\STATE
Update:
If \emph{satisfied},
\begin{align*}
p_{i,j} & =
\left\{\begin{array}{cc}
1 & \text{if $j=x_i$} \\
0 & \text{otherwise}.
\end{array}\right.
\end{align*}
If \emph{unsatisfied},
\begin{align*}
p_{i,j} =
\begin{cases}
(1-b)p_{i,j} +a/(D-1+a/b) &\text{ if } j= x_i\\
(1-b)p_{i,j} +b/(D-1+a/b) &\text{ otherwise},
\end{cases}
\end{align*}
where $a,b\in(0,1]$ are design parameters.
\ENDLOOP
\end{algorithmic}
\end{algorithm}
To understand the logic behind CFL, note that for each variable
a probability distribution over all possible variable-values is
kept. The variable's value is then selected from this distribution.
Should all the clauses that the variable participates in be satisfied
with its current value, the associated probability distribution is
updated to ensure that the variable value remains unchanged. If at
least one clause is unsatisfied, then the probability distribution
evolves by interpolating between it and a distribution that is
uniform on all values apart from on the one that is presently
generating dissatisfaction. Thus if all of a variable's clauses
were once satisfied, the algorithm retains memory of this into the
future. This has the effect that if a collection of variables are
content, they can be resistant, but not impervious, to propagation
of dissatisfaction from other variables.
CFL possesses two parameters. The parameter $b$ determines how
quickly the past is forgotten, while $a$ determines the algorithm's
aversion to a variable value should it be found to cause dissatisfaction.
Even though an instance of CFL is run for each variable, we shall show
that this completely decentralized solution is a CSP solver.
From here on we will assume that the update Step 5 is performed in
a synchronized fashion across variables. Without synchronization,
the fundamental character of the algorithm doesn't change, but the
analysis becomes more involved. Synchronization solely requires
that algorithm instances each have access to a shared sense of time
and that this can be achieved without information-sharing or other
communication between variables, i.e. between algorithm instances.
A suitable clock is, for example, available to any Internet connected
device via the Network Time Protocol (NTP).
\subsection{CFL is a decentralized CSP solver}
By construction, the only information used by the algorithm is
$\min_{m\in {\mathcal{M} }_i}\Phi_i( {\vec{x}} )$ in Step 4 and thus it satisfies the
criteria (D3) , (D4) and (D5) . That is, it only needs to know if
all clauses in which variable $i$ participates are satisfied or if
one or more are not. The CFL algorithm also satisfies the (D2)
criterion that it sticks with a solution from the first time one
is found. To see this, note that the affect of Step 5 is that if a
variable experiences success in all clauses that it participates
in it continues to select the same value with probability $1$. Thus
if all variables are simultaneously satisfied in all clauses, i.e.
if $\min_i\min_{m\in {\mathcal{M} }_i}\Phi_i( {\vec{x}} )=1$, then the same assignment
will be reselected indefinitely with probability $1$.
To establish that the CFL algorithm is a decentralized CSP solver
all that remains is to show that it meets the (D1) criterion,
that if the problem has a solution it will be found, which is dealt
with by the the following theorem. It provides an upper bound on
the distribution of the number of iterations the algorithm requires
to find a solution to any solvable CSP. It's proof exploits the
iterated function system structure of the algorithm and can be found
in the Appendix.
\begin{theorem}
\label{thm:main}
For any satisfiable CSP, with probability greater than $1-\epsilon\in(0,1)$
the number of iterations for the CFL algorithm to find a satisfying
assignment is less than
\begin{align*}
N\exp\left(\frac{N(N+1)}{2}\log(\gamma^{-1})\right)\log(\epsilon^{-1}),\\
\text{ where }
\gamma = \frac{\min(a,b)}{D-1+a/b}.
\end{align*}
For a CSP corresponding to graph coloring, a tighter bound holds
and a satisfying assignment will be found with probability greater
than $1-\epsilon$ in a number of iterations less than
\begin{align*}
N\exp(2N\log(\gamma^{-1}))\log(\epsilon^{-1}).
\end{align*}
\end{theorem}
Theorem \ref{thm:main} proves that for fixed $N$ the tail of the
distribution of the number of iterations until the first identification
of a satisfying assignment is bounded above by a geometric distribution
and so all of its moments are finite. As Theorem \ref{thm:main}
covers any arbitrary CSP that admits a solution, for any given
instance these bounds are likely to be loose. They do, however,
allow us to conclude the following corollary proving that if a
solution exists, the CFL algorithm will almost surely find it.
\begin{corollary} For any CSP that admits a satisfying assignment,
the CFL algorithm will find a satisfying assignment in almost surely
finite time.
\end{corollary}
Hence the CFL algorithm satisfies all of the criteria (D1)
(D2) (D3) (D4) (D5) , almost surely, and so is a decentralized
CSP solver.
\subsection{Parameterization}
Theorem \ref{thm:main} establishes that the CFL algorithm provably
identifies satisfying assignments for all values of its two design
parameters, $a$ and $b$. The value of $a$ determines the algorithm's
aversion to variable values for which clause failure has been
experienced. The value of $b$ impacts on the speed of convergence
of the algorithm. Optimal values of $a$ and $b$ depend upon
each problem and a performance metric. For fast convergence across
a broad range of CSPs with distinct structure, we have found that
small values of $a$ and $b$, corresponding to strong aversion to
a dissatisfying variable value and reasonably long memory,
work well. Thus for simplicity, we set $a=b$ in all experimental
examples. We use $b=0.2$ for random 3-SAT, $b=0.1$ for random 4-SAT
and $b=0.05$ for random 5-SAT. We use $b=0.1$ for our wireless
networks example, a value which we have found to yield good performance
across a range of $k$-SAT problems.
\section{Application-agnostic benchmarking}
\label{sec:sim}
\begin{table*}
\begin{center}
\begin{tabular}{|l|c|c|c|c|c|c|c|}
\hline
$k$ & 3 & 4 &5 & 7 & 10 & 20 \\
\hline
Theoretically derived bounds & & & & & & \\
\hline
$2^k \log2$ & 5.54 & 11.09 &22.18 &88.72 &709.78 & 726,817 \\
$r_{\neg exist,k}$\cite{Achlioptas05} & 4.51 & 10.23 &21.33 &87.88 &708.94 &726,817 \\
$r_{exist,k}$\cite{Achlioptas04,Kaporis06} & 3.52 & 7.91 &18.79 &84.82 &704.94 &726,809 \\
$2^k \log2-k$ & 2.54 & 7.09 &17.18 &81.72 &699.78 & 706,817 \\
$r_{poly,k}$ \cite{Frieze96,Kaporis06} & 3.52 & 5.54 &9.63 &33.23 &172.65& 95,263 \\
\hline
Estimates & & & & & & \\
\hline
$r_{\neg exist,k}$\cite{Mertens05} &4.267 & 9.93 & 21.12 & 87.79 & 708.91 & - \\
$r_{1RSB,k}$\cite{Mertens05,Krzakala07} & 4.15 & 9.38 &19.16 &62.5 & - & - \\
\hline
\end{tabular}
\caption{Current best upper/lower bounds on the phase transition
thresholds in random k-SAT.}
\label{table}
\end{center}
\end{table*}
A CSP associated with a specific networking task will possess
structure induced by its topology. Before considering practically-motivated
applications in Section \ref{sec:example}, as the CFL algorithm can
solve any CSP but its design was subject to restrictions not normally
considered it is prudent to investigate its performance as a CSP
solver in an application-agnostic setting.
We present simulation data evaluating the performance of the CFL
algorithm for random k-SAT, a well-studied class of CSPs. A k-SAT
problem is a CSP in which all variables are binary-valued and clauses
consist solely of logical disjunctions of no more than k variables.
In random k-SAT, an instance of k-SAT is generated by drawing $M$
such clauses uniformly at random \cite{papadimitriou91}. Our
investigation surprisingly reveals that the CFL algorithm is
competitive with centralized and distributed solvers on problems
of reasonable size (order 1000 variables).
Briefly, we first review current knowledge regarding
random k-SAT. The behavior of random k-SAT is known to depend
strongly on the parameter $r=M/N$ with phase
transitions and associated thresholds $r_{\neg{exist},k}$ and
$r_{{exist},k}$. If $r>r_{\neg{exist},k}$ the constraint problem
is unsatisfiable with high probability, while if $r<r_{exist,k}$
a satisfying assignments exist with high probability. Evidently,
$r_{{exist},k}\le r_{\neg{exist},k}$ and it is conjectured that
$r_{{exist},k} =r_{\neg{exist},k}$ \cite{Achlioptas04}. A simple
argument gives $r_{\neg{exist},k}\le 2^k \log 2$ and this upper
bound can be refined to obtain the values in the second row of Table
\ref{table}.
Also shown in row four of this table are estimated values for
$r_{\neg{exist},k}$ derived from statistical physics considerations.
These latter estimates are supported by experimental data
for $k=3$, e.g. \cite{NIPS04}, although there are fewer
experimental studies for $k>3$. It can be seen that the theoretical
bound for $r_{\neg{exist},k}$ and the estimated values are in good
agreement for $k\ge 5$, and that both approach $2^k \log 2$ for
large $k$. Recent mathematical results have established that
$r_{exist,k} \ge 2^k \log 2-k$ \cite{Achlioptas05} and this lower
bound can be tightened to obtain the theoretical values shown in
the third row of Table \ref{table}.
There also exists a threshold $r_{poly,k}$ below which a satisfying
assignment can be found in polynomial time with high probability.
This threshold has been the subject of much interest and the current
best analytic lower bound for $r_{poly,k}$ is also indicated in
Table \ref{table}. Statistical physics considerations have led to
the conjecture that $r_{poly,k}$ is equal to the value $r_{1RSB,k}$
at which the one-step replica symmetry breaking (1RSB) instability
occurs \cite{Mezard02}. Current best estimates for this value are
given in Table \ref{table}. For $k\ge 8$, it has been proven
analytically that the set of satisfying assignments is grouped into
widely separated clusters, which lends support to this conjecture
\cite{Mezard05}. For values of $k<8$ the situation is less clear,
with experimental evidence indicating that $r_{poly,k}$ lies above
$r_{1RSB,k}$ for $k=3,4$ and $5$ \cite{Alava08}.
\begin{figure}
\centering
\includegraphics[width=3in]{ksatvsR_median_latest2}
\caption{Median normalized stopping time of CFL algorithm vs $r$
for random 3-SAT and $N=100$. Each point is derived from at
least 1000 runs of the algorithm; algorithm terminated after $10^7$
iterations if no satisfying assignment found. For comparison,
stopping time measurements for WalkSAT taken from \cite[Fig 2a]{NIPS04}
are also indicated. (\emph{Inset}) CFL algorithm measurements for
random 4-SAT and 5-SAT. Measurements shown are for values of $r$
up to $4.20$ (3-SAT), $9.90$ (4-SAT), $21.10$ (5-SAT); close to the
current best estimates for $r_{{exist},k}$ and well above $r_{1RSB,k}$.}
\label{fig:ksatvsR}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=3in]{ksatvsN_latest}
\caption{Normalized stopping time of CFL algorithm vs $N$ and $r$
for 3-SAT and 5-SAT. Each point is derived from at least 1000
runs.}\label{fig:ksatvsN}
\end{figure}
We now consider the performance of the CFL algorithm. Fig.
\ref{fig:ksatvsR} gives median measurements of normalized
stopping time (stopping time divided by $N$) of the CFL algorithm
vs $r$=M/N when the number of variables $N=100$. Fig. \ref{fig:ksatvsN}
gives median measurements of normalized stopping time vs $N$ with
$r$ held constant. Data is shown for random k-SAT with $k=3, 4$ and
$5$. Observe two striking features from this data. Firstly, from
Fig. \ref{fig:ksatvsR} noting that a log scale is used on the
$y$-axis, the median normalized stopping time increases exponentially
with $r$ when $N$ is held constant . Secondly, from Fig.
\ref{fig:ksatvsN} the median normalized stopping time vs $N$, with
$r$ held constant, is upper bounded by a constant. That is, a
satisfying assignment is found in a time with median value that
increases no more than linearly with $N$. This linearity holds even
when $r$ is close to $r_{{exist},k}$ (data is shown in Fig.
\ref{fig:ksatvsN} for 3-SAT with $r=4$) and is of great practical
importance as it implies that with high probability the CFL algorithm
finds a satisfying assignment in polynomial time.
Comparing the performance of the CFL algorithm with the popular
WalkSAT algorithm, we note that the WalkSAT algorithm also exhibits
an exponential-like dependence of stopping time on $r$ and linearity
in $N$ \cite{NIPS04,Alekhnovich06,Alava08}. To allow comparison
in more detail, median stopping time data taken from \cite[Fig
2a]{NIPS04} is marked on Fig. \ref{fig:ksatvsR}. For $r<3.9$,
the median stopping time is similar for CFL and WalkSAT.
Above this value, however, the stopping time for WalkSAT diverges,
increasing super-exponentially in $r$. This divergence is a
feature not only of WalkSAT but also of other local search algorithms
algorithms \emph{e.g.} ChainSat \cite{Alava08}. It is not exhibited
by the Survey Propagation algorithm \cite{NIPS04,Braunstein05},
which has been the subject of considerable interest as
it creates the ability to operate close to the $r_{{exist},k}$
threshold. Observe that the CFL algorithm also does \emph{not}
exhibit divergence as $r$ approaches $r_{{exist},k}$. These
comparisons are encouraging as they indicate that the CFL algorithm
is competitive with some of the most efficient general-purpose
k-SAT algorithms currently available. It is also unexpected
as information sharing is a key component of both WalkSAT and
Survey Propagation, whereas CFL makes local decisions simultaneously
with no use whatsoever of information-sharing, raising fundamental
questions as to the role of message passing, the relationship between
information exchange and algorithm performance and, in particular,
what performance cost is imposed by constraining attention to
decentralized operation.
\section{Case study: channel allocation in 802.11 WLANs}
\label{sec:example}
Having established general properties of CFL and determined its
performance on random instances of k-SAT, we return to a problem
of the sort that motivated its introduction. We consider the
performance of the CFL algorithm in a realistic wireless network
case study. From the online database WIGLE \cite{wigle} we obtained
the locations of WiFi wireless Access Points (APs) in an approximately
$150m^2$ area at the junction of 5th Avenue and 59th Street in
Manhattan\footnote{The extracted (x,y,z) coordinate data used is
available online at \url{www.hamilton.ie/net/xyz.txt}}. This
space contains $81$ APs utilizing the IEEE 802.11 wireless
standard. It can be seen from Figure \ref{fig:meandistance}, which
plots the mean number of APs lying within distance $d$ of an AP,
that within a 15m radius an AP has on average 3 neighbours and
within a 30m radius it has on average 10 neighbours. Of the 11
channels available in the 802.11 protocol, only 3 are orthogonal.
Thus managing interference in this dense deployment is a challenging
task.
Imagine a worst-case scenario where after a power-outage all these
APs are switched back on. The aim of each AP is to select its radio
channel in such a way as to ensure that it is sufficiently different
from nearby WLANs. This can be written as a CSP where we have
$N=81$ APs and $N$ variables $x_i$ corresponding to the channel of
AP $i$, $i=1,2,...,N$. As per the 802.11 standard \cite{802.11}
and FCC regulations, each AP can select from one of 11 radio channels
in the 2.4GHz band and so the $x_i$, $i=1,2,..,N$ take values in
$ {\mathcal{D} }=\{1,2,..,11\}$. To avoid excessive interference each AP
requires that: (a) no other AP within a 5m distance operates closer
than 3 channels away; (b) no AP within a 10m distance operates
closer than 2 channels away and (c) no AP with a 30m distance
operates on the same channel. We can realize this as a CSP with
$3$ constraint clauses per AP, giving $3N$ in total,
$\Phi_m( {\vec{x}} )$, $m=1,...,3N$. With $e(i,j)$ denoting the Euclidean
distance between APs $i$ and $j$ in metres, for $m=1,...,N$
\begin{align*}
\Phi_m( {\vec{x}} )=
\begin{cases}
1 & \text{if } \min_{j:e(m,j)<5}|x_m-x_j|\ge 3 \\
0 & \text{otherwise},
\end{cases}
\end{align*}
for $m=N+1,...,2N$
\begin{align*}
\Phi_m( {\vec{x}} )=
\begin{cases}
1 & \text{if } \min_{j:e(m-N,j)<10}|x_{m-N}-x_j|\ge 2 \\
0 & \text{otherwise},
\end{cases}
\end{align*}
and for $m=2N+1,...,3N$
\begin{align*}
\Phi_m( {\vec{x}} )=
\begin{cases}
1 & \text{if } \min_{j:e(m-2N,j)<30}|x_{m-2N}-x_j|\ge 1 \\
0 & \text{otherwise},
\end{cases}
\end{align*}
The attenuation between adjacent channels is $-28$dB \cite{802.11}.
Taking the radio path loss with distance as $d^\alpha$, where $d$
is distance in meters and $\alpha=4$ the path loss exponent, then
these constraints ensure $>60$ dB attenuation between APs.
Assuming all APs use the maximum transmit power of 18dBm allowed by the 802.11 standard, this means that the SINR is greater than 20dB within a 10m radius of each AP which is sufficient to sustain a data rate of 54Mbps when the connection is line of sight and channel
noise is Gaussian \cite{802.11,goldsmith05}.
Observe that each variable $x_i$ is located at a different AP. The
APs do not belong to a single administrative domain and so security
measures such as firewalls prevent communication over the wired
network. An AP cannot rely on decoding wireless transmissions to
communicate with all of its interferers or even just to identify
them. This is because interferers may be too far away to allow their
transmissions to be decoded and yet still their aggregate transmission
power may be sufficiently powerful to create significant interference.
That is, an AP cannot know the channel selections of other APs,
condition (D3) and, moreover, cannot reliably identify or enumerate
the clauses in which it participates, condition (D4) . A decentralized
algorithm is therefore mandated.
\begin{figure}
\centering
\includegraphics[width=0.8\columnwidth]{distances_new}
\caption{Mean number of APs within distance $d$ of an AP vs distance $d$.}\label{fig:meandistance}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=3in]{5thAveSoln_new}
\caption{A satisfying assignment of radio channels. Each dot marks the location of a WiFi wireless access point and is based on measurements taken at the junction of 5th Avenue/59th
Street in Manhattan. The color of a dot indicates the radio channel. To avoid interference
between transmissions, nearby access points need to operate on radio
channels that are spaced sufficiently far apart. There are 11 radio
channels available to choose from, but the frequencies of these
channels overlap so that only 3 orthogonal/non-overlapping channels
are available; there are 81 wireless access points in total.}
\label{fig:5thAve}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=3in]{5thAveDist_new}
\caption{Log of the empirical complementary cumulative distribution
of convergence time in iterations, based on 12,000 runs of CFL
algorithm for 5th Avenue data. The median is 21 and the 95\%
percentile is 98 iterations. In current hardware one iteration can
readily be performed in under 10 seconds, leading to a median time
to convergence of less than 4 minutes.}
\label{fig:5thAveDist}
\end{figure}
Fig. \ref{fig:5thAve} shows an example satisfying assignment of
radio channels obtained by running an instance of the CFL algorithm
at all APs. The complexity of the topology generated by the physical
location of the APs and the non-uniformity of the clauses it causes
is apparent.
Fig. \ref{fig:5thAveDist} shows the measured distribution of number of
iterations required to find a satisfying assignment, whereupon the
algorithm natural halts in a decentralized fashion. The median value
is 21 iterations and the 95\% centile is 98 iterations. Note that
during this convergence period, although the network is operating
sub-optimally, it does not cease to function. In a prototype lab set-up
we have shown that a CFL update interval of less than 10 seconds
is feasible on current hardware. Thus the median time to convergence
is under 4 minutes. This is a reasonable time-frame for practical
purposes, particularly as subparts of the network are functioning
during this convergence period, and thus the CFL algorithm offers
a pragmatic solution to this difficult decentralized CSP for which
existing solvers could not be employed.
\section{Conclusions}
\label{sec:disc}
We have shown that apparently distinct problems in networking can
be placed within the framework of CSPs, where - unlike with traditional
motivating examples - the variables are co-located with devices
that may not be able to communicate. We define the criteria that
practical solvers must possess in order to be suitable for theses
problems, labeling them decentralized CSP solvers.
As existing solvers fail to meet one or more of these criteria, we
introduce a decentralized algorithm for solving CSPs. We prove that
it will almost surely find a satisfying solution if one exists. In
doing so, we generate bounds for the speed of convergence of the
algorithm. We suspect, however, that our bounds for a general CSP
are not tight and conjecture that the real bound should be closer
to the one we have for the specific case of CSPs corresponding to
graph coloring.
Given how much information decentralization is sacrificing,
surprisingly an experimental investigation of solving random k-SAT
instances suggests that the algorithm is competitive with two of
the most promising centralized k-SAT solvers, WalkSAT and Survey
Propagation, on instances of random k-SAT with order one thousand
variables. This raises the question: what is the performance cost
of decentralized operation? This is particularly pertinent as the
decentralized nature of the CFL algorithm lends itself to parallelized
computation as it has a small, fixed memory requirement per variable,
making it suitable for use in circumstances where a centralized
algorithm could also be used but would be difficult to implement
in a distributed fashion.
|
\section{Introduction}
\label{Introduction}
The ATLAS collaboration has presented in ATLAS-CONF-2010-088 {\cite{AC1088}} a
search for events with at least 3 Standard Model objects in the final state,
invariant mass above 800 GeV, and total transverse momentum above 700 GeV,
produced in 295 per nb of proton-proton collisions at centre-of-mass energy
$\sqrt{s} = 7$ TeV. \ Although not commented on in the report, the graphs of
the numbers of events against invariant mass $M_{\mathrm{{inv}}}$, which
are the right-hand graphs on pages 8 and 9 of the report, show that the
measured QCD background decreases more rapidly than the Monte Carlo over the
range $M_{\mathrm{{inv}}} = 1.1$ to 1.6 TeV, and is just half the
predicted background at 1.5 to 1.6 TeV, and there is then a bump at 1.7 to 1.9
TeV, in which the number of events per energy bin increases by a factor of
two.
The data is 3 sigma below the Monte Carlo in the $M_{\mathrm{{inv}}} =
1.5$ to 1.6 TeV bin, and although the data is only 1 sigma above the Monte
Carlo in the 1.8 to 1.9 TeV bin, the data follow a very smooth curve, with
almost no scatter, which suggests that the overall significance of the
combined dip and bump is well above 3 sigma. \ If we assume that the curve
followed by the data below and above the bump is the correct QCD background,
then the total number of events in the bump above this measured background is
8, so the bump is just under 3 sigma evidence for some sort of object in the
$s$ channel with a mass of 1.8 TeV, that produces the excess of events over
the measured background.
No update to ATLAS-CONF-2010-088 has yet been published, but since ATLAS had
collected 45 per pb of collisions by the end of 2010, the dip and bump will by
now be around 36 sigma, if the trend in the graphs has continued.
A natural interpretation of the more rapid decrease of the QCD background than
predicted by the Monte Carlo is that it represents the onset of
Dienes-\hspace{-1pt}Dudas-Gherghetta or Arkani-Hamed-Cohen-Georgi accelerated
unification of the Standard Model coupling constants {\cite{DDG1, DDG2, ACG}},
which begins at about a third of the energy at which compact extra dimensions
accessible to the Standard Model gauge fields start to become visible.
In this article I shall present rough estimates to show that the bump could
arise from about $10^{30}$ approximately degenerate
Kaluza-Klein states of the $d = 11$ supergravity multiplet
in the $s$ channel, which according to a conjecture by Brooks, reported in
\cite{KMST}, on the
lowest nonzero eigenvalue of the Laplace-Beltrami operator on generic compact
hyperbolic manifolds of large intrinsic volume, might arise from
compactification of $d = 11$ supergravity on a manifold with a compact
hyperbolic Cartesian factor of intrinsic volume around $10^{34}$, and
curvature radius around an inverse TeV.
In section \ref{The bulk}, on page \pageref{The bulk}, I shall look for
solutions of the quantum-corrected Einstein equations of $d = 11$ supergravity
{\cite{CJS}} that have 4 flat extended dimensions and a compact 7-manifold
with one or more compact hyperbolic factors, in the presence of magnetic
4-form fluxes whose bilinears are on average covariantly constant. \ The
necessary flux-dependent terms in the dimension 8 local term in the $d = 11$
quantum effective action are not yet known, and I shall use a simple guess for
them for the case when the flux is covariantly constant, which on reduction to
10 dimensions with NS 3-form flux agrees with the Kehagias-Partouche
conjecture {\cite{KP}}, supported by recent calculations by Richards
{\cite{Richards 2}}, up to correction terms that contain factors that
occur in the classical Einstein equation in 10 dimensions.
Using this action solutions are proved impossible for the $\bar{H}^7$,
$\bar{H}^5 \times \bar{H}^2$, and $\bar{H}^4 \times \bar{H}^3$ cases even when
5-branes are added whose world-volumes lie along the 4 extended dimensions and
wrap 2-cycles of the compact manifold $\mathcal{M}^7$, and various types of
search have found no solutions in the $\bar{H}^5 \times S^2$, $\bar{H}^4
\times S^3$, $\bar{H}^3 \times S^4$, and $\bar{H}^3 \times S^2 \times S^2$
cases. \ Here $\bar{H}^n$ denotes a smooth compact quotient of $n$-dimensional
hyperbolic space $H^n$ by a freely-acting discrete subgroup of
$\mathrm{{SO}} \left( n, 1 \right)$, which is the isometry group of
$H^n$. \ However at this level of approximation, where only the leading
quantum corrections to the classical action of $d = 11$ supergravity are
included, the existence of solutions is affected by field redefinitions, which
should have no effect when all orders of perturbation theory are included. \
An admissible field redefinition results in the existence of a solution in the
$\bar{H}^7$ case, and probably also in the other cases. \ A Ho\v{r}ava-Witten
boundary to accommodate the strong/electroweak Standard Model fields is
introduced in section \ref{The HW boundary}, on page \pageref{The HW
boundary}.
The motivation for considering such solutions of quantum-corrected $d = 11$
supergravity is that the space-time we live in is approximately flat up to
distances larger, by a factor of around $10^{61}$, than the radius of
curvature that would be expected from combining the Standard Model (SM) of the
strong/electroweak interactions with Einstein gravity in $3 + 1$ dimensions. \
Rigid extended objects such as the jib of a crane, a wing or the hull of an
aeroplane, or an oil tanker, which are much longer in one dimension than in
their other two dimensions, have a structure that makes the fullest possible
use of their extension in their two short dimensions, in order to stiffen them
in their long dimension. \ These examples suggest that the large size and
flatness of the space-time we live in are supported by structures that make
use of the extension of the universe in small additional spatial dimensions.
Compact hyperbolic manifolds $\bar{H}^n$ with $n \geq 3$ have an ideal
structure for stiffening the extended dimensions in this way, because when the
metric on them is locally symmetric, and their curvature is fixed, they are
completely rigid {\cite{Mostow 1, Mostow 2, Prasad, Thurston, Gromov
Hyperbolic}}. \ Their shape and size is fixed by their topology, and they can
be arbitrarily large. \ In particular, the volume $\bar{V}$ of a compact
hyperbolic manifold $\bar{H}^n$ when its sectional curvature is equal to $-
1$, which I shall call its \emph{intrinsic volume}, is fixed by its
topology, and can be arbitrarily large.
Supergravity in $10 + 1$ dimensions is the unique quantum field theory that
includes gravity, and whose gauge invariances force the cosmological constant
to vanish in its defining dimension {\cite{CJS, Nicolai Townsend van
Nieuwenhuizen, Sagnotti Tomaras, Bautier Deser Henneaux Seminara}}, and $10 +
1$ is the largest number of dimensions in which supergravity can exist if
there is just one time dimension and flat space-time is required to be a
solution of the classical field equations {\cite{Nahm}}. \ Thus a solution of
the quantum corrected field equations of $d = 11$ supergravity on an almost
flat $\mathbf{R}^4$ times a compact hyperbolic 7-manifold $\bar{H}^7$ of
large intrinsic volume, or failing that a compact 7-manifold $\mathcal{M}^7$
with at least one Cartesian factor $\bar{H}^n$ with $n \geq 3$ and large
intrinsic volume, might provide a reasonable candidate model for the
space-time we live in.
$d = 11$ supergravity can only exist on manifolds whose topology allows them
to have a spin structure. \ Such manifolds are called \emph{spin
manifolds}. \ All orientable manifolds of dimensions $n \leq 3$ are spin
manifolds, and Cartesian products of spin manifolds are spin manifolds
{\cite{Michelson Lawson}}. \ A manifold $\mathcal{M}$ is spin if and only if
its second Stiefel-Whitney class $w_2 \left( \mathcal{M} \right) \in H^2
\left( \mathcal{M}; \mathbf{Z}_2 \right)$ vanishes {\cite{Hirzebruch et
al}}, where $H^2 \left( \mathcal{M}; \mathbf{Z}_2 \right)$ is the second
cohomology group with $\mathbf{Z}_2$ coefficients {\cite{Milnor Stasheff}}.
\ Thus in the absence of additional information, the chance that an
$n$-dimensional orientable manifold $\mathcal{M}^n$ with $n \geq 4$ is spin is
$2^{- B_{2, \mathbf{Z}_2}}$, where $B_{2, \mathbf{Z}_2}$, the second Betti
number with $\mathbf{Z}_2$ coefficients, is the number of $\mathbf{Z}_2$
factors in $H^2 \left( \mathcal{M}; \mathbf{Z}_2 \right)$. \ The sum of the
Betti numbers of an $n$-dimensional compact hyperbolic manifold $\bar{H}^n$
with intrinsic volume $\bar{V}$ is bounded above by $b \bar{V}$ for arbitrary
coefficients, where $b$ is a constant that depends only on $n$ and the
coefficient ring {\cite{Gromov Volume Bounded Cohomology, Gromov Volume and
Bounded Cohomology}}, and the number
of topologically distinct $\bar{H}^n$ with intrinsic volume $\bar{V}$ less
than a fixed value $V$ is finite for fixed $n \geq 4$ {\cite{Wang}}, but grows
with $V$ as $V^{cV}$ for sufficiently large $V$, where $c > 0$ is a constant
that depends only on $n$ {\cite{Burger Gelander Lubotzky Mozes}}. \ The Davis
manifold, which is an $\bar{H}^4$, is spin {\cite{Davis manifold, Ratcliffe
Tschantz Davis manifold}}, so we may reasonably expect that for all $n \geq
2$, there are spin $\bar{H}^n$, and that for $n \geq 4$, their number will
grow with $V$ as $V^{cV}$ for sufficiently large $V$, while for $n = 3$, their
number is already infinite for $V \geq 2.03$ {\cite{Cao Meyerhoff, Thurston}}.
Supergravity does not couple to any matter fields in $10 + 1$ smooth
dimensions, but if the $10 + 1$ dimensional space-time has a $9 + 1$
dimensional smooth ``boundary'', which could perhaps better be described as a
flexible mirror, because all the fields on one side of the mirror are exactly
copied, up to sign, on the other side of the mirror, then supergravity in the
$10 + 1$ dimensional ``bulk'' couples to a supersymmetric Yang-Mills multiplet
{\cite{Wess Zumino, Gliozzi Scherk Olive}}, with gauge group $E_8$, on the $9
+ 1$ dimensional ``boundary''. \ The Yang-Mills multiplet is adjacent to the
flexible mirror, but infinitesimally displaced from it, so that it has its own
reflection infinitesimally on the other side of the mirror {\cite{Lu}}. \ This
is called Ho\v{r}ava-Witten (HW) theory {\cite{Horava Witten 1, Horava Witten 2,
Moss 1, Moss 2, Moss 3, Moss 4}}, and I shall assume that we live on an HW
boundary, while the graviton is part of the supergravity multiplet in the $10
+ 1$ dimensional bulk.
An $n$-dimensional compact hyperbolic space $\bar{H}^n$, $n \geq 2$, is a
quotient of $n$-dimen-sional hyperbolic space, $H^n$, by a discrete subgroup
$\Gamma$ of the symmetry group $\mathrm{{SO}} \left( n, 1 \right)$ of
$H^n$, such that $\Gamma$ acts freely on $H^n$, which means that no element of
$\Gamma$ other than the identity leaves any point of $H^n$ invariant. \ This
is equivalent to the requirement that $\Gamma$ have no nontrivial finite
subgroup, which in the language of group theory is expressed by saying that
$\Gamma$ has no torsion. \ The construction of a large family of $\bar{H}^n$,
called arithmetic manifolds {\cite{Borel Harish Chandra}}, is reviewed in
{\cite{Witte Morris}} and section 3 of {\cite{CCHT}}. \ Non-arithmetic
manifolds were constructed from suitable pairs of arithmetic manifolds for all
$n \geq 2$ in {\cite{Gromov Piatetski Shapiro}}. \ It is the rapid growth with
intrinsic volume $\bar{V}$ of the number of topologically distinct finite
coverings of the non-arithmetic manifolds that results in the $V^{cV}$ growth
of the number of topologically distinct $\bar{H}^n$ with $\bar{V} \leq V$
{\cite{Burger Gelander Lubotzky Mozes}}.
For the special case of $n = 3$, an infinite number of topologically distinct
$\bar{H}^3$ can be constructed from each finite-volume cusped hyperbolic
3-manifold $\hat{H}^3$ by applying an operation called Dehn filling to each
cusp, and each $\bar{H}^3$ constructed in this way has $\bar{V} < \hat{V}$,
where $\hat{V}$ is the intrinsic volume of the initial $\hat{H}^3$
{\cite{Thurston}}. \ A large number of $\bar{H}^3$ of small $\bar{V}$ have
been constructed and catalogued in this way, and their properties can be
studied using computer programs {\cite{SnapPea, SnapPy, Snap}}. \ The
Cartesian product of two $\bar{H}^3$ of small $\bar{V}$ might be a possible
topology for the compact Cartesian factor of the HW boundary that accommodates
the Standard Model fields. \ A \emph{cusp} of a finite-volume hyperbolic
$n$-manifold $\hat{H}^n$ is a region of the manifold whose topology is the
Cartesian product of an infinite half-line and a flat $\left( n - 1
\right)$-manifold, for example an $\left( n - 1 \right)$-torus, such that the
area of the cross section decreases so rapidly along the half-line that the
volume of the cusp is finite. \ The number of cusps is always finite.
The $n$-dimensional hyperbolic space $H^n$ can be realized as the hypersurface
$t^2 - \vec{x}^2 = 1$, $t > 0$, in $n + 1$ dimensional Minkowski space. \
Using spherical polar coordinates for $\vec{x}$, the equation reduces to $t^2
- r^2 = 1$, $t > 0$, so we can choose $t = \mathrm{\cosh} \rho$, $r =
\mathrm{\sinh} \rho$, and $\rho$ is then the geodesic distance of the point
$\left( \rho, \theta_1, \ldots, \theta_{n - 1} \right)$ from the origin
{\cite{Costa}}. \ The Riemann tensor of $H^n$ has the form:
\begin{equation}
\label{Riemann tensor of Hn} R_{i j k l} = g_{i l} g_{j k} - g_{i k} g_{j
l},
\end{equation}
where $g_{i j}$ is the metric, which means that $H^n$ has constant sectional
curvature equal to $- 1$. \ The Ricci tensor of $H^n$ is then $R_{i j} = -
\left( n - 1 \right) g_{i j}$. \ The Riemann tensor of a $d$-dimensional space
or space-time, with metric $G_{I J}$, is defined in general by:
\begin{equation}
\label{Riemann tensor} R_{I J} \, \!^K \, \!_L = \partial_I \Gamma_J \, \!^K
\, \!_L - \partial_J \Gamma_I \, \!^K \, \!_L + \Gamma_I \, \!^K \, \!_M
\Gamma_J \, \!^M \, \!_L - \Gamma_J \, \!^K \, \!_M \Gamma_I \, \!^M \,
\!_L,
\end{equation}
and the Ricci tensor and scalar are defined by $R_{I J} = R_{K I} \, \!^K \,
\!_J$, and $R = G^{I J} R_{I J}$. \ The standard Christoffel connection is
$\Gamma_I \, \!^J \, \!_K = \frac{1}{2} G^{J L} \left( \partial_I G_{L K} +
\partial_K G_{L I} - \partial_L G_{I K} \right)$.
I shall use units such that $\hbar = c = 1$. \ The gravitational coupling
constant in 11 dimensions is $\kappa_{11}$. \ The metric is mostly $+$. \
Coordinate indices $I, J, K, \ldots$ run over all 11 dimensions, and
coordinate indices $\mu, \nu, \sigma, \ldots$ are tangent to the four observed
space-time dimensions. \ The bosonic fields of supergravity in 11 dimensions
are the graviton $G_{I J}$, and a 3-index antisymmetric tensor Abelian gauge
field $C_{I J K}$, called a 3-form gauge field, with 4-form field strength
$H_{I J K L} = \partial_I C_{J K L} - \partial_J C_{K L I} + \partial_K C_{L I
J} - \partial_L C_{I J K}$.
The bosonic part of the classical action of $d = 11$ supergravity is
{\cite{CJS}}:
\begin{equation}
\label{CJS action} S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{CJS}}} = \frac{1}{2 \kappa_{11}^2}
\int_{\mathcal{B}} d^{11} xe \left( R - \frac{1}{48} H_{I J K L} H^{I J K L}
- \frac{1}{144^2} \epsilon^{I_1 \ldots I_{11}}_{11} C_{I_1 I_2 I_3} H_{I_4
\ldots I_7} H_{I_8 \ldots I_{11}} \right),
\end{equation}
where $\mathcal{B}$ means the bulk. \ The sign of the Einstein term follows
from the Riemann and Ricci tensor conventions stated above, and the
normalization of the 4-form kinetic term, and the definition of $H_{I J K L}$
given above, are chosen to agree with {\cite{Hyakutake Ogushi 2}}. \ $e =
\sqrt{- G}$ is the determinant of the vielbein $e_{I \hat{J}}$, where $G$ is
the determinant of the metric $G_{I J}$, and the antisymmetric tensor
$\epsilon^{I_1 \ldots I_{11}}_{11}$ is related to the $\mathrm{{SO}}
\left( 10, 1 \right)$ invariant tensor $\epsilon^{\hat{I}_1 \ldots
\hat{I}_{11}}_{11}$, with components $0, \pm 1$, by $\epsilon^{I_1 \ldots
I_{11}}_{11} = e^{I_1} \, \!_{\hat{J}_1} \ldots e^{I_{11}} \,
\!_{\hat{J}_{11}} \epsilon^{\hat{J}_1 \ldots \hat{J}_{11}}_{11}$. \ Hatted
indices are local Lorentz indices. \ $C_{I J K}$ here is related to $C_{I J
K}$ of {\cite{Horava Witten 2, Moss 1, Moss 2, Moss 3, Moss 4}} by $C_{I J K
\left( \mathrm{{here}} \right)} = 6 \sqrt{2} C_{I J K \left(
\mathrm{{HW}} \right)}$, and $H_{I J K L}$ is related to $G_{I J K L}$ of
{\cite{Horava Witten 2, Moss 1, Moss 2, Moss 3, Moss 4}} by $H_{I J K L} =
\sqrt{2} G_{I J K L}$.
I assume the unperturbed metric in 11 dimensions has the form:
\begin{equation}
\label{metric ansatz} ds_{11}^2 = G_{IJ} dx^I dx^J = a \left( x^C \right)^2
\eta_{\mu \nu} dx^{\mu} dx^{\nu} + h_{AB} dx^A dx^B
\end{equation}
where $\eta_{\mu \nu} = \mathrm{{diag}} \left( - 1, 1, 1, 1 \right)$ is
the metric on $\left( 3 + 1 \right)$-dimensional Minkowski space, and $h_{AB}$
is the metric on the compact 7-dimensional manifold-with-boundary
$\mathcal{M}^7$. \ Coordinate indices $A, B, C, \ldots$ are tangential to
$\mathcal{M}^7$. \ The ``warp factor'' $a \left( x^C \right)$ is equal to a
constant value $A$ everywhere on $\mathcal{M}^7$ except in the immediate
vicinity of the HW boundary, and it is equal to 1 on the HW boundary. \ I
shall show in section \ref{The HW boundary}, starting on page \pageref{The HW
boundary}, that a consistent semiclassical treatment of the boundary region is
possible if $A$ is $< 1$ but comparable to 1, for example $A = 0.7$. \ I
assume that $\mathcal{M}^7$ has an $n$-dimensional compact hyperbolic
Cartesian factor $\breve{H}^n$ of large intrinsic volume $\breve{V}_n$, which
has one boundary of small intrinsic area $\bar{A}_{n - 1}$, where $\bar{A}_{n
- 1}$ is defined using the metric of sectional curvature $- 1$ on
$\breve{H}^n$, and that the 10-dimensional HW boundary is the Cartesian
product of that boundary, and the remaining Cartesian factor $\mathcal{M}^{7 -
n}$ of $\mathcal{M}^7$, and the 4 extended dimensions. \ $\breve{H}^n$ is
constructed by cutting a complete $n$-dimensional compact hyperbolic manifold
$\bar{H}^n$ of large intrinsic volume $\bar{V}^n$ along a small $\left( n - 1
\right)$-cycle that separates it into two disconnected parts. \ The smaller of
the two parts is discarded, and the other becomes $\breve{H}^n$.
The Einstein action in the 4 extended dimensions has the form:
\begin{equation}
\label{Einstein action} S_{\mathrm{{Ein}}} = \frac{1}{16 \pi G_N} \int
d^4 x \sqrt{- g} g^{\mu \nu} R_{\mu \nu} \left( g \right),
\end{equation}
where $g_{\mu \nu}$ differs from $\eta_{\mu \nu}$ by a small perturbation,
that depends on the coordinates $x^{\sigma}$ on the 4 extended dimensions, but
not on the coordinates $x^A$ on $\mathcal{M}^7$. \ $G_N$ is Newton's constant,
with the value {\cite{PDG}}:
\begin{equation}
\label{Newtons constant} G_N = 6.7087 \times 10^{- 33}
\textrm{{ TeV}}^{- 2},
\end{equation}
so that $\sqrt{G_N} = 8.1907 \times 10^{- 17}$ TeV$^{- 1} \hspace{3pt} =
\hspace{3pt} 1.6160 \times 10^{- 35}$ metres. \ Comparing with (\ref{CJS
action}), and noting that $R_{I J} \, \!^K \, \!_L$ and hence $R_{I J}$ are
unaltered by rescaling the metric by a constant factor, so that $\sqrt{- G}
G^{\mu \nu} R_{\mu \nu} \left( G \right) = A^4 \sqrt{- g} \sqrt{h}
\frac{1}{A^2} g^{\mu \nu} R_{\mu \nu} \left( g \right)$ everywhere on
$\mathcal{M}^7$ except in the immediate vicinity of the HW boundary, where
$G_{I J}$ here represents the metric obtained from (\ref{metric ansatz}) by
replacing $\eta_{\mu \nu}$ by $g_{\mu \nu}$, we find, in the approximation of
neglecting the volume of the region where $a \left( x^C \right)$ differs
appreciably from $A$, that the volume $V_7 \equiv \int_{\mathcal{M}^7} d^7 x^C
\sqrt{h}$ of $\mathcal{M}^7$ is given by {\cite{ADD1, AADD, ADD2, RS1}}:
\begin{equation}
\label{V7 in terms of GN} \frac{A^2 V_7}{2 \kappa^2_{11}} = \frac{1}{16 \pi
G_N}
\end{equation}
The experimental limits on the gravitational coupling constant in $D$
dimensions are expressed in terms of a mass $M_D$, such that for $D = 11$,
$M_{11} = \left( 2 \pi \right)^{7 / 9} \kappa^{- 2 / 9}_{11} = 4.1764
\kappa^{- 2 / 9}_{11}$. \ The latest limit on $M_{11}$ for 7 flat extra
dimensions, from {\cite{Franceschini et al}}, including bounds from
non-observation of effects of virtual graviton exchange and graviton emission
at the LHC, is that $M_{11}$ is larger than about 1.5 TeV, so that $\kappa^{-
2 / 9}_{11}$ is larger than about 0.36 TeV. \ The limit might be less
stringent for hyperbolic large extra dimensions, since the light Kaluza-Klein
modes will be far fewer or absent. \ The dimension $n$ of the large intrinsic
volume hyperbolic Cartesian factor $\breve{H}^n$ of $\mathcal{M}^7$ might be
less than 7, but the lower limit on $M_D$ found in {\cite{Franceschini et al}}
is about 1.5 TeV for all $D \geq 6$. \ If $\kappa^{- 2 / 9}_{11}$ was about
0.36 TeV and $A$ was 1, then from (\ref{Newtons constant}) and (\ref{V7 in
terms of GN}), $V_7$ would be about $5.8 \times 10^{34}$ TeV$^{- 7}$.
The classical field equations of $d = 11$ supergravity do not have any
solutions with 4 flat extended dimensions if $\mathcal{M}^7$ has a compact
hyperbolic factor, for any arrangement of 4-form fluxes such that the flux
bilinears are on average covariantly constant, but in section \ref{The bulk} I
shall show that when the leading quantum corrections to the classical action
are taken into account, together with the freedom to make redefinitions of the
fields of the form $G_{I J}, C_{I J K} \rightarrow G_{I J} + X_{I J}, C_{I J
K} + Y_{I J K}$, where $X_{I J}$ and $Y_{I J K}$ are polynomials in the fields
and their derivatives, with sufficiently small coefficients, then there is a
solution with an almost flat $\mathbf{R}^4$ times a compact hyperbolic
7-manifold $\bar{H}^7$ of large intrinsic volume, with magnetic 4-form fluxes
wrapping 4-cycles of the $\bar{H}^7$. \ Field redefinitions of this type are
like a change of coordinates in ``field space'', so they do not change the
physical content of the theory. \ They do not alter the $S$-matrix
{\cite{Diagrammar, Lee Les Houches, Tseytlin Field Redefinitions}}, and when
the leading higher-order corrections to the classical CJS action are
calculated by requiring anomaly cancellation on 5-branes {\cite{Duff Liu
Minasian, Witten 5 branes, Freed Harvey Minasian Moore, Bilal Metzger,
Harvey}} and in Ho\v{r}ava-Witten theory {\cite{Horava Witten 2, de Alwis, Conrad,
Faux Lust Ovrut, Lu, Bilal Derendinger Sauser, Harmark, Bilal Metzger,
Meissner Olechowski, Moss 3}}, and supersymmetry {\cite{Hyakutake Ogushi 1,
Hyakutake Ogushi 2, Hyakutake, Hyakutake 2,
Metsaev}}, the result is only determined up to the freedom to add
arbitrary linear combinations of terms that vanish when the classical field
equations are satisfied, because the coefficients of such linear combinations
of terms can be adjusted arbitrarily by small amounts, by field redefinitions
of this type.
The leading quantum corrections are local terms that depend on the curvature
of $\mathcal{M}^7$ but not directly on its topology, so the curvature radius
$B$ of the $\bar{H}^7$ is around $\kappa^{2 / 9}_{11}$. \ The value of $B$
depends on the field redefinition parameter, but using the ``principle of
minimal sensitivity'' advocated for dealing with the renormalization scheme
dependence of low order calculations in perturbative QCD {\cite{Stephenson}},
the best value of $B$ is around $0.43 \kappa^{2 / 9}_{11}$. \ Thus if
$\kappa^{- 2 / 9}_{11}$ was about 0.36 TeV, $B$ would be around 1.2 TeV$^{-
1}$, so if $A$ was 1, so that $V_7$ was about $5.8 \times 10^{34}$ TeV$^{-
7}$, the intrinsic volume $\bar{V}$ of the $\bar{H}^7$ would be around $1.6
\times 10^{34}$. \ The $\mathbf{R}^4$ could be exactly flat but for the
quantization of the 4-form fluxes {\cite{Rohm Witten, Witten flux
quantization}}, and for an $\bar{H}^7$ of intrinsic volume $\bar{V}^7 \sim
10^{34}$, the de Sitter radius of the $\mathbf{R}^4$ can easily be as large
as the observed de Sitter radius, which is $16.0$ Gyr $= \hspace{1pt}
\hspace{1pt} 1.51 \times 10^{26}$ metres $= \hspace{1pt} \hspace{1pt}
\hspace{2pt} 0.94 \times 10^{61} \sqrt{G_N}$ {\cite{HST, 0302209 Spergel et
al, de Sitter radius}}.
The diameter $L$ of a compact manifold is by definition the maximum over all
pairs of points of the manifold of the shortest geodesic distance between
them, and I shall call $\bar{L}$, the diameter of a compact hyperbolic
manifold $\bar{H}^n$ when its sectional curvature is equal to $- 1$, its
\emph{intrinsic diameter}. \ If $\bar{H}^n$ is reasonably isotropic, in
the sense that it has a fundamental domain in $H^n$ that is approximately
spherical, then using spherical polar coordinates on $H^n$ as above, the
intrinsic volume $\bar{V}_n$ of $\bar{H}^n$ is approximately related to its
intrinsic diameter $\bar{L}$ by
\begin{equation}
\label{V bar n} \bar{V}_n \simeq S_{n - 1} \int^{\bar{L} / 2}_0
\mathrm{\sinh}^{n - 1} \rho d \rho,
\end{equation}
where $S_{n - 1}$ is the area of the $\left( n - 1 \right)$-sphere of unit
radius. \ Thus if $\bar{H}^n$ is reasonably isotropic and has large intrinsic
volume $\bar{V}_n$, the relation between $\bar{V}_n$ and $\bar{L}$ is
approximately
\begin{equation}
\label{V bar n for large volume} \bar{V}_n \simeq \frac{S_{n - 1}}{2^{n - 1}
\left( n - 1 \right)} e^{\left( n - 1 \right) \frac{\bar{L}}{2}} .
\end{equation}
Thus for reasonably isotropic $\bar{H}^7$ with volume $V_7 \simeq 5.8 \times
10^{34}$ TeV$^{- 7}$ and curvature radius $B$ around 1.2 TeV$^{- 1}$, so that
$\bar{V}_7 \simeq 1.6 \times 10^{34}$, we find from $S_6 = \frac{16}{15}
\pi^3$ that the intrinsic diameter of $\bar{H}^7$ is $\bar{L} \simeq 27$, so
its actual diameter is $L \simeq 32 \textrm{{ TeV}}^{- 1}$.
On a flat Riemannian manifold of diameter $L$, the lowest non-zero eigenvalue
of the negative of the Laplace-Beltrami operator $\Delta \equiv
\frac{1}{\sqrt{g}} \partial_i \left( \sqrt{g} g^{i j} \partial_j \right)$,
which is the generally covariant form of the Laplacian, is typically $\left(
\frac{2 \pi}{L} \right)^2$. \ However there is some evidence that the lowest
non-zero eigenvalue $\lambda_1$ of $- \Delta$ on a reasonably isotropic
compact hyperbolic manifold $\bar{H}^n$ with sectional curvature $- 1$ and
large diameter $\bar{L}$ will be $\sim 1$ rather than $\sim
\frac{1}{\bar{L}^2}$ {\cite{KMST}}. \ Firstly, the spectrum of $- \Delta$ on
$H^n$, which is completely continuous, starts at $\lambda = \frac{\left( n - 1
\right)^2}{4}$, rather than 0 {\cite{Yau Isoperimetric, Agmon}}. \ Secondly,
for $n = 2$, Brooks and Makover have constructed very large families of
compact Riemann surfaces with sectional curvature $- 1$ and large genus, all
of which have $\lambda_1 > \frac{3}{16}$ {\cite{Brooks Makover 1, Brooks
Makover 2, Brooks Makover 3, Brooks Makover 4}}.
And thirdly, we can use Cheeger's inequality {\cite{Cheeger, Cheeger
Wikipedia}}, which states that for a compact Riemannian manifold
$\mathcal{M}$, the lowest non-zero eigenvalue $\lambda_1$ of $- \Delta$ on
$\mathcal{M}$ is bounded below by:
\begin{equation}
\label{Cheegers inequality} \lambda_1 \left( \mathcal{M} \right) \geq
\frac{h^2 \left( \mathcal{M} \right)}{4},
\end{equation}
where $h \left( \mathcal{M} \right)$, called the Cheeger isoperimetric
constant of $\mathcal{M}$, is defined by:
\begin{equation}
\label{Cheeger constant} h \left( \mathcal{M} \right) \equiv
\textrm{min}_{\mathcal{N}} \left( \frac{A \left( \mathcal{N}
\right)}{\mathrm{min} \left( V \left( \mathcal{M}_1 \right), V \left(
\mathcal{M}_2 \right) \right)} \right),
\end{equation}
where $\mathcal{N}$ denotes a hypersurface in $\mathcal{M}$ that divides
$\mathcal{M}$ into two disconnected parts $\mathcal{M}_1$ and $\mathcal{M}_2$,
$V \left( \mathcal{M}_i \right)$ is the volume of $\mathcal{M}_i$, and $A
\left( \mathcal{N} \right)$ is the area of $\mathcal{N}$, or in other words,
if $\mathcal{M}$ is $n$-dimensional, the $\left( n - 1 \right)$-volume of
$\mathcal{N}$, calculated in the given metric on $\mathcal{M}$.
To apply Cheeger's inequality to a large volume compact hyperbolic manifold
$\bar{H}^n$ with sectional curvature $- 1$ that is reasonably isotropic, in
the sense that it has an approximately spherical fundamental domain, we note
that a totally geodesic hypersurface through the centre of the approximately
spherical fundamental domain will not, in general, divide the manifold into
two disconnected parts, because the surface of the fundamental domain consists
of a large number of $\left( n - 1 \right)$-dimensional polyhedral faces, each
of which is identified with another face typically roughly on the opposite
side of the fundamental domain. \ To minimize the value of $\frac{A \left(
\mathcal{N} \right)}{\mathrm{\min} \left( V \left( \mathcal{M}_1 \right), V
\left( \mathcal{M}_2 \right) \right)}$, we first look for minimal area
topologically non-trivial hypersurfaces $\mathcal{N}$ that divide $\bar{H}^n$
into two disconnected parts of approximately equal volume, and we see that a
typical such $\mathcal{N}$ will consist of one totally geodesic hypersurface
through the centre of the fundamental domain, plus polyhedral pieces that
cover exactly half the surface of the fundamental domain. \ Using spherical
polar coordinates as before, we see that for $\bar{H}^n$ of large intrinsic
volume and diameter $\bar{L}$, the area of the part of $\mathcal{N}$ that
covers half the surface of the fundamental domain is
\begin{equation}
\label{A of mathcal N} A \left( \mathcal{N}_{\mathrm{{half}}
\,\,\mathrm{{sphere}}} \right) \simeq \frac{S_{n - 1}}{2^n}
e^{\left( n - 1 \right) \frac{\bar{L}}{2}} \simeq \frac{\left( n - 1
\right)}{2} \bar{V}_n,
\end{equation}
and the area of the part of $\mathcal{N}$ that is a totally geodesic
hypersurface through the origin is negligible in comparison to this,
suggesting that the Cheeger constant of $\bar{H}^n$ is $h \left( \bar{H}^n
\right) \simeq n - 1$. \ As a check on this, we consider spherical
$\mathcal{N}$ of radius $\rho$ centred at the centre of the fundamental
domain. \ For all $\rho \gg 1$ the area of such an $\mathcal{N}$ is $n - 1$
times the volume it encloses, so for all $\rho \gg 1$, $\rho \leq \bar{L} -
\frac{2 \mathrm{\ln} 2}{n - 1}$, the value of $\frac{A \left( \mathcal{N}
\right)}{\mathrm{\min} \left( V \left( \mathcal{M}_1 \right), V \left(
\mathcal{M}_2 \right) \right)}$ is again $n - 1$. \ Thus it looks likely that
for all reasonably isotropic compact hyperbolic manifolds $\bar{H}^n$ with
sectional curvature $- 1$ and large intrinsic volume, the lowest nonzero
eigenvalue $\lambda_1$ of $- \Delta$ on $\bar{H}^n$ will by Cheeger's
inequality be bounded below by:
\begin{equation}
\label{lower bound on lambda 1} \lambda_1 \geq \frac{\left( n - 1
\right)^2}{4}
\end{equation}
This is the value of $\lambda$ at which the spectrum starts on $H^n$
\cite{Donnelly}, and the
lower bound on $\lambda$ for $H^n$ could be obtained from Cheeger's inequality
using spheres of radius $\rho \gg 1$ as above, if Cheeger's inequality could
be applied to $H^n$.
Furthermore, considering now $\bar{H}^n$ with sectional curvature $- 1$ and
large intrinsic volume that depart somewhat from being reasonably isotropic,
there is for $n = 3$ a family of manifolds called Bianchi manifolds of
arbitrarily large but finite intrinsic volume, that depart to the greatest
possible extent from being reasonably isotropic, namely they have a finite
number of cusps as described just before (\ref{Riemann tensor of Hn}) above,
for which $\lambda_1$ has been bounded below by $21 / 25 = 0.84$ and is
conjectured to be equal to 1 {\cite{Sarnak, Luo, Rudnick Sarnak, Orlando
Park}}. \ The diameter of a cusped hyperbolic $n$-manifold $\hat{H}^n$ of
finite volume and sectional curvature $- 1$ is infinite, and each cusp
contributes a continuous part to the spectrum of $- \Delta$ on $\hat{H}^n$. \
However just as for $H^n$, the continuous part of the spectrum starts at
$\lambda = \frac{\left( n - 1 \right)^2}{4}$, rather than 0 {\cite{Mazzeo
Phillips, Donnelly Xavier}}. \ For $n = 3$ the Dehn filling construction
produces from any cusped $\hat{H}^n$ of finite volume an infinite sequence of
compact $\bar{H}_i^n$ that converge to $\hat{H}^n$ as $i \rightarrow \infty$
in the sense that the intrinsic volumes converge, and the $\bar{H}_i^n$ look
almost identical to $\hat{H}^n$ except for regions that move further and
further out along the cusps of $\hat{H}^n$ as $i \rightarrow \infty$
{\cite{Thurston}}, and the eigenvalues of $- \Delta$ below the bottom of the
continuous spectrum of $\hat{H}^n$ are limits of the eigenvalues of the
$\bar{H}_i^n$ {\cite{Colbois Courtois 1, Colbois Courtois 2}}. \ Thus for $n =
3$ there are infinite sequences of compact $\bar{H}_i^n$ with sectional
curvature $- 1$ that converge to fixed cusped Bianchi manifolds, which have
arbitrarily large but finite intrinsic volumes, and these $\bar{H}_i^n$ have
$\lambda_1$ converging to a number that is $\geq 0.84$ and conjectured to be
1, while their diameters tend to $\infty$ and their intrinsic volumes tend to
fixed finite values, that can, however, be arbitrarily large.
So it seems possible, at least for $n = 2$ and $n = 3$, that a compact
hyperbolic manifold $\bar{H}^n$ with sectional curvature $- 1$ and arbitrarily
large intrinsic volume will have $\lambda_1 \simeq \frac{\left( n - 1
\right)^2}{4}$ even if it departs substantially from being isotropic. \ This
is in agreement with a conjecture by Brooks reported in {\cite{KMST}}. \ I
shall now assume that this is the case, which means that classically, the
lightest massive spin 2 Kaluza-Klein graviton would have mass $\sim \kappa^{-
2 / 9}_{11}$ and thus around a TeV in the geometry considered above.
In contrast with the spectrum of $- \Delta$, the spectrum of the Dirac operator
on the $n$-dimensional hyperbolic space $H^n$ is the whole real line
{\cite{Bunke, Bar}}. \ Thus in view of the correspondence between the spectrum
of $- \Delta$ on $H^n$ and on a compact hyperbolic manifold $\bar{H}^n$ of
large intrinsic volume $\bar{V}$ and intrinsic diameter $\bar{L}$ as
considered above, it seems likely that classically, the lightest gravitino
modes would have masses $\sim \frac{2 \pi}{\bar{L} B}$, and thus around 0.2
TeV in the geometry considered above, but just as in the case of large flat
extra dimensions, their effects would not show up at the LHC below energies
around $M_{11}$, which is about 1.5 TeV in this example.
The $\mathbf{R}_{\mathrm{{flat}}}^4 \times \bar{H}^7$ solution of the
quantum-corrected field equations of $d = 11$ supergravity found in section
\ref{The bulk} depends essentially on the leading quantum corrections to the
classical action, so there will be corresponding corrections $\sim \kappa^{- 2
/ 9}_{11}$ to the masses of all the Kaluza-Klein modes. \ I shall assume
provisionally that the effect of these corrections is to increase the masses
of the Kaluza-Klein modes. \ This is particularly significant for the 3-form
gauge field $C_{I J K}$, since a compact hyperbolic manifold $\bar{H}^n$ of
large intrinsic volume $\bar{V}$ has Betti numbers $B_1, \ldots, B_{n - 1}$
that in a particular family of examples are bounded below by constants times
powers $>0$ and $<1$ of $\bar{V}$ {\cite{Xue}}, and for $n = 7$ these would
lead classically to $B_1$ massless 2-form Abelian gauge fields, each equivalent
in $3 + 1$ dimensions to a massless scalar \cite{Cremmer Scherk, Kalb Ramond},
$B_2$ massless Abelian vector gauge fields, and $B_3$ massless scalars. \ There
would also be $B_1$ additional massless Abelian vector gauge fields arising
from
$G_{I J}$. \ I shall assume provisionally that all these classically massless
modes aquire masses $\sim \kappa^{- 2 / 9}_{11}$ from the leading quantum
corrections.
If we write the eigenvalues of $- \Delta$ on an $n$-dimensional compact
manifold $\mathcal{M}^n$, with volume $V_n$, as $k^2$, then for large
$\bar{k}$, the number of modes with $k \leq \bar{k}$ is approximately given by
the Weyl asymptotic formula as
\begin{equation}
\label{number of modes up to k bar} \frac{V_n}{\left( 2 \pi \right)^n} S_{n
- 1} \frac{\bar{k}^n}{n},
\end{equation}
which corresponds to a density of states $\frac{V_n}{\left( 2 \pi \right)^n}$
in momentum space. \ In numerical studies of the spectrum of $- \Delta$ on
compact hyperbolic 3-manifolds of small intrinsic volume, Inoue found that the
Weyl asymptotic formula is actually approximately valid down to $\bar{k}^2
\simeq \lambda_1$, the lowest non-zero eigenvalue, so that if $\lambda_1$
occurs at a larger value of $\bar{k}$ than would be expected from the Weyl
formula, there is a degeneracy or approximate degeneracy of eigenvalues near
$\lambda_1$, that restores agreement with the Weyl formula for $\bar{k}^2$
above $\lambda_1$ {\cite{Inoue}}.
It seems reasonable to expect that a similar approximate degeneracy of
eigenvalues near $\lambda_1$, that restores agreement with the Weyl asymptotic
formula (\ref{number of modes up to k bar}) for $\bar{k}^2 > \lambda_1$, will
also occur for compact hyperbolic manifolds of large intrinsic volume, since
the number of modes up to $\bar{k} = \sqrt{\lambda_1}$ given by (\ref{number
of modes up to k bar}) is already large, comparable to the intrinsic volume
$\bar{V}_n$. \ I shall provisionally assume that this happens, which means that
for the
geometry described above, we expect in the region of $10^{34}$ approximately
degenerate Kaluza-Klein modes of the gravitational multiplet, close to a mass
$\simeq \frac{n - 1}{2 B} = \frac{3}{B} \simeq 2.5$ TeV.
In partial support of this hypothesis,
every $\bar{H}^n$ has pairs of finite covers of arbitrarily large volume
ratio, whose sets of eigenvalues of $-\Delta$, ignoring multiplicities, are
identical \cite{Leininger et al}, so since the Weyl asymptotic formula
(\ref{number of modes up to k bar}) is certainly valid for sufficiently large
$\bar{k} = \sqrt{\lambda}$,
every $\bar{H}^n$ has finite covers whose eigenvalues have arbitrarily large
multiplicities, for sufficiently large $\bar{k}$.
I shall now provisionally assume that the bump at 1.7 to 1.9 TeV seen in the
right-hand
graphs on pages 8 and 9 of ATLAS-CONF-2010-088 arises from the production and
decay of these approximately $10^{34}$ approximately degenerate gravitational
Kaluza-Klein modes in the $s$ channel, and give rough estimates to test
this interpretation. \ I shall show in section
\ref{The HW boundary}, starting on page \pageref{The HW boundary}, that a mode
on $\bar{H}^7$ of intrinsic mass $k$ has mass $\frac{Ak}{B}$ as measured on
the HW boundary, where $B \simeq 0.43 \kappa^{2 / 9}_{11}$ is the curvature
radius of the $\bar{H}^7$, and $A$ is the value of the ``warp factor'' $a
\left( x^C \right)$ everywhere on the $\bar{H}^7$ except in the immediate
vicinity of the HW boundary, as described after (\ref{metric ansatz}). \ Thus
if the bump centred at 1.8 TeV corresponds to approximately degenerate modes
of intrinsic mass $\frac{n - 1}{2} = 3$ on the $\bar{H}^7$, and
$\kappa^{- 2 / 9}_{11}$ is at its current lower bound for 2 or more flat extra
dimensions, $\kappa^{- 2 / 9}_{11} \simeq 0.36$ TeV \cite{Franceschini et al},
as in the example above, then $A =
\frac{1.8 \hspace{0.3em} \mathrm{{TeV}}}{2.5 \hspace{0.3em}
\mathrm{{TeV}}} = 0.72$. \ This then revises $V_7$ from the value $5.8
\times 10^{34}$ TeV$^{- 7}$ calculated from (\ref{V7 in terms of GN}) assuming
$A = 1$, to $1.1 \times 10^{35}$ TeV$^{- 7}$, and the intrinsic volume
$\bar{V}_7$ is revised to $3.1 \times 10^{34}$.
For an order of magnitude estimate, I shall treat each Kaluza-Klein mode of
the graviton multiplet, with mass around 1.8 TeV, as an independent scalar
particle $\varphi$. \ Then the total cross section for the process $gg
\rightarrow \varphi \rightarrow u \bar{u}$, for example, where $g$ represents
a gluon, is given near 1.8 TeV by the Breit-Wigner cross-section
{\cite{Perkins}}, and in order of magnitude is:
\begin{equation}
\label{Breit Wigner cross section} \sigma \left( g g \rightarrow \varphi
\rightarrow u \bar{u} \right) \sim \frac{\Gamma_{g g} \Gamma_{u
\bar{u}}}{E^2} \cdot \frac{1}{\left( \left( E - M_{\varphi} \right)^2 +
\Gamma^2 / 4 \right)},
\end{equation}
where $\Gamma_{gg}$ and $\Gamma_{u \bar{u}}$ are the partial widths for
$\varphi$ to decay to $g g$ and $u \bar{u}$ respectively, $E$ is the invariant
mass of the $g g$ system, $M_{\varphi} \simeq 1.8$ TeV is the mass of
$\varphi$, and $\Gamma$ is the total width of $\varphi$. \ The partial widths
are estimated in order of magnitude by {\cite{Giudice Rattazzi Wells}}:
\begin{equation}
\label{partial widths} \Gamma_{g g} \sim \Gamma_{g g g} \sim \ldots \sim
\Gamma_{u \bar{u}} \sim \Gamma_{g u \bar{u}} \sim \ldots \sim
\Gamma_{\nu_{\tau} \bar{\nu}_{\tau}} \sim M_{\varphi}^3 G_N \simeq 3.9
\times 10^{- 32} \textrm{{TeV}}
\end{equation}
The only factor in the cross-section (\ref{Breit Wigner cross section}) that
varies significantly with energy over the energy range of interest is the
final factor, and this arises from the squared magnitude of the
energy-dependent factor:
\begin{equation}
\label{energy dependent factor} \frac{1}{E - M_{\varphi} + i \Gamma / 2}
\end{equation}
in the corresponding amplitude. \ The Kaluza-Klein graviton is not present in
the final state, so we have to sum the amplitude over the contributions of all
the different Kaluza-Klein graviton modes, all of which give the same
contribution to the amplitude, apart from slight differences in their masses.
The number of bosonic modes in the $d = 11$ supergravity multiplet is $128$,
so the number of approximately degenerate bosonic Kaluza-Klein states of mass
$\simeq 1.8$ TeV, which is the number of states up to 1.8 TeV as given by the
Weyl asymptotic formula (\ref{number of modes up to k bar}), is around:
\begin{equation}
\label{number of degenerate KK modes} 128 \frac{\bar{V}_7}{7 \left( 2 \pi
\right)^7} S_6 \,\, 3^7 \simeq 1.1 \times 10^{35}
\end{equation}
where I used $\bar{V}_7 \simeq 3.1 \times 10^{34}$ from just before
(\ref{Breit Wigner cross section}), and $S_6 = \frac{16}{15} \pi^3$, and
(\ref{lower bound on lambda 1}) with $ n=7 $. \ I shall
assume that these states are distributed over a smooth peak in the density of
states from about 1.7 TeV to 1.9 TeV.
The total width $\Gamma$ of $\varphi$ is the product of the partial width
(\ref{partial widths}) and the number of Standard Model states, which is 36,
so $\Gamma$ is extremely small. \ We can replace the sum over the individual
Kaluza-Klein modes by an integral over their masses {\cite{Franceschini et
al}}, and to a first approximation, when convoluted with a smooth density of
states, the energy-dependent factor (\ref{energy dependent factor}) is
effectively $\simeq - i \pi \delta \left( E - M \right)$, because the $i
\Gamma$ means that the integration path has to go around the singularity in
the lower half of the complex $M$ plane. \ Choosing the integration path to be
along the real axis except for a small semicircle centred at $M = E$, the
contributions from the real axis approximately cancel, and the semicircle
gives $- i \pi$ times the density of states evaluated at $M = E$. \ Thus after
summing the amplitude over the Kaluza-Klein modes, the energy-dependent factor
in the amplitude is approximately replaced by:
\begin{equation}
\label{energy dependent factor summed over modes} - 1.1 \times 10^{35} i \pi
f \left( E \right),
\end{equation}
where $f \left( E \right)$ has a smooth peak from about 1.7 TeV to 1.9 TeV,
and $\int f \left( E \right) dE = 1$. \ Thus the cross-section is:
\begin{equation}
\label{cross section summed over KK modes} \sigma \left( g g \rightarrow
\varphi \rightarrow u \bar{u} \right) \sim \frac{\Gamma_{g g} \Gamma_{u
\bar{u}}}{E^2} \times 1.2 \times 10^{70} \pi^2 f \left( E \right)^2 \sim 5.6
\times 10^7 f \left( E \right)^2 .
\end{equation}
A monoenergetic high energy beam of protons with $N$ protons of energy $E$ per
unit area per unit time is equivalent to a beam of partons, such that the
number of $u$ quarks per unit area per unit time with energy between $xE$ and
$\left( x + dx \right) E$ is $f_u \left( x \right) dx$, and similarly for the
other types of parton, where $f_p \left( x \right)$, $p = u, d, g, \bar{u},
\bar{d}, \ldots$ are the parton distribution functions (PDFs). \ The PDFs
evolve logarithmically with $Q^2$, the square of the momentum transferred in a
scattering process, and for a rough estimate I shall use the gluon
distribution $f_g \left( x \right)$ from the plot {\cite{MSTW}} with $Q^2 =
10^4$ GeV$^2$. \ To produce $\varphi$ with mass 1.8 TeV at rest with 3.5 TeV
per proton beam, each gluon needs $x = 0.26$. \ If the 4-momenta of the beams
are $\left( P, 0, 0, P \right)$ and $\left( P, 0, 0, - P \right)$ and the
momentum fractions of the gluons are $x_1$ and $x_2$, then their Mandelstam
$s$ is $P^2 \left( \left( x_1 + x_2 \right)^2 - \left( x_1 - x_2 \right)^2
\right) = 4 P^2 x_1 x_2$. \ From the plot we find that $f_g \left( x \right)
\simeq 0.060 x^{- 2.17}$ for $0.05 \leq x \leq 0.2$, but substantially smaller
than this for $x \geq 0.3$. \ Then the total cross section to produce a
$\varphi$ within the bump is roughly:
\[ \int^{0.3}_0 dx_1 \int^{0.3}_0 dx_2 0.060^2 \left( x_1 x_2
\right)^{- 2.17} \times 5.6 \times 10^7 f \left( 2 P \sqrt{x_1 x_2}
\right)^2 \]
\begin{equation}
\label{cross section for bump} \simeq \frac{2.0 \times 10^5}{\left( 0.2
\textrm{{ TeV}} \right)^2} \times \int^{0.3}_{0.23} \frac{dx_1}{x_1}
\int^{0.27}_{0.24} \frac{2 xdx}{x^{4.34}} \simeq 3.0 \times 10^6
\textrm{{ nb}},
\end{equation}
where I approximated $f \left( E \right)$ as a constant from 1.7 TeV to 1.9
TeV and 0 outside this interval. \ Multiplying by $4 \pi$, and by 75 for the
number of Standard Model helicity states, and by $\frac{1}{8}$ for the
probability that the two initial gluons can form a colour singlet, and by
$\frac{1}{4}$ for the average over the helicity states of the initial gluons,
gives a final estimate of $8.8 \times 10^7$ nb for the total cross section to
produce a $\varphi$ within the bump that decays to two Standard Model objects,
if $\kappa^{- 2 / 9}_{11} \simeq 0.36$ TeV, and all the bosonic modes of the
$d = 11$ supergravity multiplet that according to the Weyl asymptotic formula
(\ref{number of modes up to k bar}) would have masses up to 1.8 TeV, are in
fact approximately degenerate, with masses in the range 1.7 TeV to 1.9 TeV.
To reduce the background, the ATLAS measurement selected events with at least
3 Standard Model objects in the final state, and total transverse momentum
above 700 GeV. \ However the partial widths for $\varphi$ to decay to 2, 3, or
4 Standard Model states are all of the same order of magnitude (\ref{partial
widths}). \ And for an order of magnitude estimate, the decay of $\varphi$ is
effectively isotropic, and the 700 GeV lower limit on the total tranverse
momentum will not result in missing a high percentage of the decays of the 1.8
TeV $\varphi$ into 3 or 4 Standard Model states. \ The number of different
types of 3 object final states is roughly 10 times the number of different
types of 2 object final states, and omitting this factor could very roughly
compensate for neglecting the effect of the 700 GeV lower limit on the total
tranverse momentum. \ Thus we find an estimate of $10^8$ nb for the total
cross section to produce a $\varphi$ within the bump that is counted as an
event in the ATLAS measurement, if $\kappa^{- 2 / 9}_{11} \simeq 0.36$ TeV,
and all the bosonic modes of the $d = 11$ supergravity multiplet that
according to (\ref{number of modes up to k bar}) would have masses up to 1.8
TeV, are in fact approximately degenerate, with masses in the range 1.7 TeV to
1.9 TeV. \ So in 295 per nb of proton-proton collisions there should be about
$3 \times 10^{10}$ events above background in the bump. \ If we assume that the
curve followed by the data below and above the bump is the correct QCD
background, then the total number of events in the bump above this measured
background is 8.
Thus the estimate of the number of modes in the bump, (\ref{number of
degenerate KK modes}), is too large by a factor of around 60000,
so the correct number is around $1.8 \times 10^{30} $. \ From
(\ref{V7 in terms of GN}) and (\ref{number of degenerate KK modes})
and the relation $V_7 = B^7 \bar{V}_7$, where
$B \simeq 0.43 \kappa^{2 / 9}_{11}$
is the curvature radius of the $\bar{H}^7$, and the
relation $\frac{3 A}{B} = 1.8$ TeV, where $A$ is the value of the ``warp
factor'' $a \left( x^C \right)$ everywhere on the $\bar{H}^7$ except in the
immediate vicinity of the HW boundary, as described after (\ref{metric
ansatz}), we see that the intrinsic volume
$\bar{V}_7$ of $\bar{H}_7$, which determines the number of modes in the bump
if all the modes expected from (\ref{number of modes up to k bar}) up to the
expected intrinsic $\lambda_1 = 3^2$ are approximately degenerate at 1.8 TeV,
is independent of $\kappa^{2 / 9}_{11}$, so the discrepancy cannot be
corrected by reducing $\kappa^{2 / 9}_{11}$.
However the expected number of
modes in the bump would be reduced by a factor of $3^{- 7} = \frac{1}{2187}$
to around $ 5.0 \times 10^{31} $ if for some reason the intrinsic $\lambda_1$
was 1 rather than $3^2$ as expected from Cheeger's inequality. \ This might
happen if the modes in the bump originate from modes of the 3-form $C_{I J K}$
that are 2-forms on $\bar{H}^7$ and vectors along the extended dimensions,
because for $p < \frac{n - 1}{2}$, $\lambda_1$ for a $p$-form on $H^n$ is
$\frac{\left( n - 1 - 2 p \right)^2}{4}$ {\cite{Donnelly Xavier}}.
In this case the relation $\frac{3 A}{B} = 1.8$ TeV would be modified to
$\frac{A}{B} = 1.8$ TeV, so since $A \leq 1$ by section \ref{The HW boundary},
starting on page \pageref{The HW boundary}, we would find from $B \simeq 0.43
\kappa^{2 / 9}_{11}$ that $\kappa^{- 2 / 9}_{11} \geq 0.78
\textrm{{ TeV}}$, hence $M_{11} \geq 3.3$ TeV.
If this is the correct interpretation, then the number of modes would actually
be reduced still further, possibly to around the correct value, because the
factor 128 in (\ref{number of degenerate KK modes}) would be replaced by a
smaller value, representing the fact that the modes in the bump represent only
modes of $C_{I J K}$ that are 2-forms on $\bar{H}^7$ and vectors along the
extended dimensions. \ In this case, there might be a second bump at around $2
\times 1.8 \textrm{{ TeV}} = 3.6 \textrm{{ TeV}}$ from modes of
$C_{I J K}$ that are 1-forms on $\bar{H}^7$ and 2-form Abelian gauge fields,
equivalent in 4 dimensions to scalars \cite{Cremmer Scherk, Kalb Ramond}, along
the extended dimensions, and from
modes of the metric $G_{I J}$ that are vectors on $\bar{H}^7$ and vectors along
the extended dimensions, and a
third bump at around $3 \times 1.8 \textrm{{ TeV}} = 5.4
\textrm{{ TeV}}$ from the KK modes of the graviton that are scalars on
$\bar{H}^7$ and traceless symmetric tensors along the extended dimensions.
However the position is made less clear by the fact that, as discussed just
before (\ref{number of modes up to k bar}), on page \pageref{number of modes
up to k bar}, I had to assume that the classically massless modes of
$ C_{I J K} $ and $G_{I J}$
resulting from the Hodge - de Rham harmonic forms on $\bar{H}^7$ all acquire
masses $ \sim \kappa_{11}^{-2/9} $ from the leading quantum corrections to
the classical CJS action (\ref{CJS action}).
If the intrinsic $\lambda_1$ is $3^2$, as assumed on the basis of Cheeger's
bound for scalars on $\bar{H}^7$ in the initial estimate above of the expected
number of events above background in the bump, then we would have to conclude
that just a fraction around $\frac{1}{60000}$ of
the bosonic gravitational KK modes expected from (\ref{number of modes up to k
bar}) are approximately degenerate at 1.8 TeV.
An alternative hypothesis is that notwithstanding the rapid restoration of
agreement between the spectral staircase
and the Weyl asymptotic formula (\ref{number of modes up to k bar}) found for
small intrinsic volume $\bar{H}^3$'s in \cite{Inoue}, and the possibility of
arbitrarily large degeneracies of eigenvalues of $-\Delta$ for large intrinsic
volume $\bar{H}^n$'s indicated by \cite{Leininger et al}, the restoration of
agreement with (\ref{number of modes up to k bar}) for $\bar{H}^7$ occurs over
a wider range of energies than 0.2 TeV, and the modes responsible for the bump
from 1.7 to 1.9 TeV are instead the Hodge - de Rham harmonic 1-forms, 2-forms,
and 3-forms on $\bar{H}^7$, which have all acquired approximately equal masses
$\sim \kappa_{11}^{-2/9}$ from the $(DH)^2 R^2$ and $H^2 R^3$ terms in the
leading quantum corrections to the CJS action (\ref{CJS action}).
The number of linearly independent Hodge - de Rham harmonic $p$-forms, $0 \leq
p \leq n$, on a compact orientable $n$-manifold, is equal to the $p$th Betti
number with integer coefficients, $B_p$, which is called the $p$th Betti
number. The sum of all the Betti numbers of an $\bar{H}^n$ of intrinsic
volume $\bar{V}_n$ is bounded above by $c \bar{V}_n$, where $c$ is a constant
that depends only on $n$ \cite{Gromov Volume Bounded Cohomology, Gromov Volume
and Bounded Cohomology}. The $p$th Betti number $B_p$ of finite coverings of
an $\bar{H}^n$ grows linearly with the inrinsic volume of the covering if and
only if there are square-integrable harmonic $p$-forms on the covering space
$H^n$ \cite{Lueck, Clair
Whyte}, which by \cite{Donnelly Xavier} means if and only if $n$ is even and
$p=\frac{n}{2}$. In the other cases, the rate of growth of $B_p$ with the
intrinsic volume of the covering is bounded above by a rate of growth that is
strictly less than linear, by an amount that depends on the rate of growth of
the
\emph{injectivity radius} of the covering. The injectivity radius
$\mathrm{Inj}\left(\mathcal{M},x\right)$ at a point
$x$ of a Riemannian manifold $\mathcal{M}$ is the largest radius for which the
map from the tangent space at $x$, to $\mathcal{M}$, defined by mapping each
tangent vector $v$ at $x$ to the point reached by starting at $x$ and going out
a distance $|v|$ along the geodesic starting in the direction $v$ from $x$, is
a diffeomorphism, and $\mathrm{Inj}\left(\mathcal{M}\right)$ is the minimum of
$\mathrm{Inj}\left(\mathcal{M},x\right)$ over all the points $x$ of
$\mathcal{M}$ \cite{Wikipedia injectivity radius}.
If $\bar{H}^n$ is a compact hyperbolic $n$-manifold, and $X$ is a finite regular covering manifold of $\bar{H}^n$ with intrinsic volume
$\mathrm{Vol}\left(X \right)$ and intrinsic injectivity radius
$\mathrm{Inj}\left(X \right)$, then by Theorem 0.3 of
\cite{Clair Whyte}, there are constants $C > 0$ and $\beta_p > 0$ such that:
\begin{itemize}
\item For $n$ odd:
\begin{itemize}
\item If $p \neq \frac{n \pm 1}{2}$, then $B_p\left(X \right) \leq C
\displaystyle{
\frac{\mathrm{Vol}\left(X \right)}{e^{\beta_p \mathrm{Inj}\left(X \right)}}}$.
\item For $p = \frac{n \pm 1}{2}$, $B_p\left(X \right) \leq C \displaystyle{
\frac{\mathrm{Vol}\left(X \right) \cdot \mathrm{log}\,\,\mathrm{Inj}\left(X
\right)}{\mathrm{Inj}\left(X \right)}}$.
\end{itemize}
\item For $n$ even:
\begin{itemize}
\item If $p \neq \frac{n}{2}$, then $B_p\left(X \right) \leq C
\displaystyle{
\frac{\mathrm{Vol}\left(X \right)}{e^{\beta_p \mathrm{Inj}\left(X \right)}}}$.
\end{itemize}
\end{itemize}
By the generalized Gauss-Bonnet theorem {\cite{Allendoerfer Weil, Chern}}, the
Euler
characteristic, or Euler number, $\chi \left( \mathcal{M}^{2 q} \right) \equiv
\sum^{2 q}_{p = 0} \left( - \right)^p B_p$, of an arbitrary smooth $2
q$-manifold $\mathcal{M}^{2 q}$, is given by:
\begin{equation}
\label{generalized Gauss Bonnet theorem} \chi \left( \mathcal{M}^{2 q}
\right) = \frac{1}{\left( 8 \pi \right)^q q!} \int_{\mathcal{M}^{2 q}} d^{2
q} x \sqrt{g} \epsilon_{j_1 \ldots j_{2 q}} \epsilon^{i_1 \ldots i_{2 q}}
R_{i_1 i_2} \, \!^{j_1 j_2} \ldots R_{i_{2 q - 1} i_{2 q}} \, \!^{j_{2 q -
1} j_{2 q}} .
\end{equation}
Thus for a $2 q$-dimensional compact hyperbolic manifold $\bar{H}^{2 q}$ with
sectional curvature $- 1$:
\begin{equation}
\label{Euler number of H bar 2q} \chi \left( \bar{H}^{2 q} \right) = \left(
- \right)^q \frac{\left( 2 q \right)\! !}{\left( 2 \pi \right)^q 2^q q!}
\bar{V}_{2 q} = \left( - \right)^q \frac{\left(2 q - 1\right)\! !\! !}{\left( 2
\pi \right)^q} \bar{V}_{2 q} = \left( - \right)^q \frac{2}{S_{2 q}}
\bar{V}_{2 q},
\end{equation}
where $S_{2 q} = \frac{2 \left( 2 \pi \right)^q}{\left( 2 q - 1 \right) !!}$
is the $2 q$-volume of the $2 q$-sphere of unit radius. \ In particular, $\chi
\left( \bar{H}^6 \right) = - \frac{15}{8 \pi^3} \bar{V}_6 \simeq - 0.060
\bar{V}_6$, $\chi \left( \bar{H}^4 \right) = \frac{3}{4 \pi^2} \bar{V}_4
\simeq 0.076 \bar{V}_4$, and $\chi \left( \bar{H}^2 \right) = - \frac{1}{2
\pi} \bar{V}_2 \simeq - 0.159 \bar{V}_2$. \ Thus since by {\cite{Lueck, Clair
Whyte}} and {\cite{Donnelly Xavier}}, $B_p \left( X \right)$ for a finite
regular covering $X$ of $\bar{H}^{2 q}$ can grow linearly with the intrinsic
volume $\mathrm{{Vol}} \left( X \right)$ if and only if $p = q$, $\left(
- \right)^q \frac{B_q \left( X \right)}{\chi \left( X \right)} \rightarrow 1$
as $\mathrm{{Vol}} \left( X \right) \rightarrow \infty$, hence $\frac{B_q
\left( X \right)}{\mathrm{{Vol}} \left( X \right)} \rightarrow
\frac{2}{S_{2 q}}$ as $\mathrm{{Vol}} \left( X \right) \rightarrow
\infty$.
The relatively strong suppression of $B_p\left(X \right)$ relative to
$\mathrm{Vol}\left(X \right)$ for $p\neq\frac{n\pm 1}{2}$ when $n$ is odd and
for $p\neq\frac{n}{2}$ when $n$ is even arises from the fact that in these
cases, the spectrum of the Hodge - de Rham $-\Delta$ for $p$-forms on $H^n$
has a gap \cite{Donnelly Xavier}. For $p = \frac{n\pm 1}{2}$ when $n$ is odd
the spectrum of $-\Delta$ on $H^n$ has no gap, but there are no
square-integrable harmonic $p$-forms on $H^n$, while for $p = \frac{n}{2}$
when $n$ is even, there are an infinite number of linearly independent
square-integrable harmonic $p$-forms on $H^n$ \cite{Donnelly Xavier}.
For finite coverings $X$ of $\bar{H}^n$ that are reasonably isotropic, in the
sense that they have a fundamental domain that is approximately spherical, we
might expect that the intrinsic injectivity radius $\mathrm{Inj}\left(X
\right)$, which is half the intrinsic length of the shortest closed geodesic on
$X$, is
roughly a fixed fraction of the intrinsic diameter of $X$, in which case for
for $p\neq\frac{n\pm 1}{2}$ when $n$ is odd and for $p\neq\frac{n}{2}$ when $n$
is even, the above bounds on $B_p\left(X \right)$ become upper bounds by
constant multiples of powers strictly less than 1 of $\mathrm{Vol}\left(X
\right)$. In the special case of $\bar{H}^n$ whose fundamental groups are
arithmetic discrete subgroups $\Gamma$ of $\mathrm{SO}\left(n,1\right)$
\cite{Borel Harish Chandra}, and finite coverings $X$ of $\bar{H}^n$ whose
fundamental groups are congruence subgroups of $\Gamma$, which are roughly the
subgroups obtained by Selberg's Lemma \cite{Selberg}, reviewed in subsection
3.1.2 of \cite{CCHT}, there are lower bounds on $B_p\left(X \right)$ of this
form \cite{Xue}.
If we assume that the above results for finite coverings $X$ of a fixed
$\bar{H}^n$ roughly describe the dependence of the Betti numbers $B_p\left(
\bar{H}^n\right)$ on the intrinsic volume $\bar{V}$ of $\bar{H}^n$ for typical
$\bar{H}^n$ of large $\bar{V}$, at least for reasonably isotropic $\bar{H}^n$,
then for $p\neq\frac{n\pm 1}{2}$ when $n$ is odd and for $p\neq\frac{n}{2}$
when $n$ is even, $B_p\left(\bar{H}^n\right)$ is roughly a constant times a
power $<1$ of $\bar{V}$, while for $p = \frac{n\pm 1}{2}$ when $n$ is odd,
$B_p\left(\bar{H}^n\right)$ is roughly a constant times $\frac{\bar{V}}{
\mathrm{ln}\,\bar{V}}$, and for $p = \frac{n}{2}$ when $n$ is even, $B_p\left(
\bar{H}^n \right)$ is roughly $\frac{2}{S_n}\bar{V}$.
Thus if the $\bar{H}^n$ Cartesian factor of
$\mathcal{M}^7$ of large $\bar{V}$ had $n$ even, the number of modes in the bump
according to the alternative hypothesis would probably be too large by a factor
of at least 1000. For the $\bar{H}^7$ case, the number of modes would
be about right if $B_3$ and $B_4$ were around $\frac{1}{200}\,\frac{\bar{V}}{
\mathrm{ln}\,\bar{V}}$.
If this is the correct hypothesis, then the bump will be spin 0 if the
compact hyperbolic factor of large intrinsic volume is 7-dimensional, and a
mixture of spins 0 and 1 if it is 5-dimensional or
3-dimensional. This could be tested by plotting the energies of the decay
products in 3-body decays of candidate bump particles, in the reconstructed
rest frame of the candidate bump particle, on a Dalitz plot \cite{Dalitz,
Fabri, Perkins}.
Confirmation that the bump particles decay democratically to all Standard
Model particles, and are thus gravity-like, but have only spin 0, would
provide evidence for the existence of the 3-form gauge field $C_{I J K}$ of
$d = 11$ supergravity \cite{CJS}, because in the $\bar{H}^7$ case, the bump
particles arise only from $C_{I J K}$.
\vspace{0.3cm}
\section{Ho\v{r}ava-Witten theory}
\label{Horava Witten theory}
I shall use Moss's improved version of Ho\v{r}ava-Witten theory {\cite{Moss 1,
Moss 2, Moss 3, Moss 4}}. \ In the region of the HW boundary, the coordinates
$x^I$ have the form $\left( \tilde{x}^U, y \right)$, where indices $U, V, W,
\ldots$ are tangential to a family of hypersurfaces foliating the $\left( 10 +
1 \right)$-dimensional manifold-with-boundary, one of these hypersurfaces
coinciding with the boundary, and $y$ takes a constant value on each of these
hypersurfaces, with the value of $y$ distinguishing the hypersurfaces. \ $y$
takes the value $y_1$ on the boundary, and $y > y_1$ in the bulk. \ The symbol
$y$ is also used as the
coordinate index for the $y$ coordinate. \ The bosonic part of the
semiclassical action is:
\begin{equation}
\label{Horava Witten action} S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{HW}}} = S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{CJS}}} + S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{GH}}} + S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{YM}}}
\end{equation}
The supergravity term is (\ref{CJS action}), on page \pageref{CJS action}.
The Gibbons-Hawking term is:
\begin{equation}
\label{Gibbons Hawking term} S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{GH}}} = \frac{1}{\kappa^2_{11}} \int_{\beta} d^{10}
\tilde{x} \tilde{e} K
\end{equation}
Here $\beta$ denotes the $9 + 1$ dimensional boundary at $y = y_1$, and on the
boundary, $\tilde{e} = \sqrt{- \tilde{G}}$ denotes the square root of minus
the determinant of the induced metric $\tilde{G}_{U V}$, which is obtained
from $G_{I J}$ by dropping the row and column with an index $y$. \ $K = G^{I
J} K_{I J}$ is the trace of the extrinsic curvature $K_{I J} \equiv \left(
\delta_I \, \!^K - n_I n^K \right)\! \left( \delta_J \, \!^L - n_J n^L \right)
D_K n_L$, where $n_I$ is the outward unit normal, and $D_K$ is the covariant
derivative. \ Thus since $n_I$ is a scalar factor times $\partial_I y$, the
only nonvanishing component of $n_I$ is $n_y = - \frac{1}{\sqrt{G^{y y}}}$, so
only the $D_U n_V = - \Gamma_U \, \!^y \, \!_V n_y$ components of $D_K n_L$
contribute to $K_{I J}$, so $K_{I J}$ is symmetric, and $K =
\frac{1}{\sqrt{G^{y y}}} \left( G^{U V} - \frac{G^{U y} G^{V y}}{G^{y y}}
\right) \Gamma_U \, \!^y \, \!_V$.
The Yang-Mills term is:
\begin{equation}
\label{Yang Mills term} S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{YM}}} = - \frac{1}{16 \pi \kappa^2_{11}} \left(
\frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \int_{\beta} d^{10} \tilde{x}
\tilde{e} \left( \frac{1}{30} \mathrm{{tr}} F_{U V} F^{U V} -
\frac{1}{2} \bar{R}_{U V \hat{W} \hat{X}} \bar{R}^{U V \hat{W} \hat{X}}
\right)
\end{equation}
Here $F_{U V} = \partial_U A_V - \partial_V A_U + i \left[ A_U, A_V \right]$
is the field strength of an $E_8$ Yang-Mills gauge field $A_U =
T^{\mathcal{A}} A_U^{\mathcal{A}}$ localized on the boundary. \
$\mathrm{{tr}}$ means the ordinary trace, not the modified trace used by
HW. \ Indices $\mathcal{A}, \mathcal{B}, \ldots$ run over the 248 generators
of $E_8$, and the hermitian generators $T^{\mathcal{A}}$ in the
fundamental/adjoint of $E_8$ satisfy $\mathrm{{tr}} T^{\mathcal{A}}
T^{\mathcal{B}} = 30 \delta^{\mathcal{A}\mathcal{B}}$. \ In the
$\mathrm{{SO}} \left( 16 \right)$ basis for $E_8$, the $T^{\mathcal{A}}$
are $- \frac{1}{2} i$ times the generators in Appendix 6.A of {\cite{Green
Schwarz Witten}} or subsection 2.1 of {\cite{CCHT}}, and in the
$\mathrm{{SU}} \left( 9 \right)$ basis for $E_8$, the $T^{\mathcal{A}}$
are the generators in subsection 5.2 of {\cite{CCHT}}. \ The coefficient of the
first term in
(\ref{Yang Mills term}) is fixed by anomaly cancellation {\cite{Horava Witten
2, de Alwis, Conrad, Faux Lust Ovrut, Lu, Bilal Derendinger Sauser, Harmark,
Bilal Metzger, Meissner Olechowski, Moss 3}} and has the value found by
Conrad {\cite{Conrad}}, which is slightly different from the original value
found by HW. \ The $\bar{R}_{U V \hat{W} \hat{X}} \bar{R}^{U V \hat{W}
\hat{X}}$ term was derived by Moss {\cite{Moss 4}}, with:
\begin{equation}
\label{R bar} \bar{R}_{U V} \, \!^{\hat{W}} \, \!_{\hat{X}} = \partial_U
\bar{\omega}_V \, \!^{\hat{W}} \, \!_{\hat{X}} - \partial_V \bar{\omega}_U
\, \!^{\hat{W}} \, \!_{\hat{X}} + \bar{\omega}_U \, \!^{\hat{W}} \,
\!_{\hat{Y}} \bar{\omega}_V \, \!^{\hat{Y}} \, \!_{\hat{X}} - \bar{\omega}_V
\, \!^{\hat{W}} \, \!_{\hat{Y}} \bar{\omega}_U \, \!^{\hat{Y}} \,
\!_{\hat{X}},
\end{equation}
where
\begin{equation}
\label{omega bar} \bar{\omega}_{U \hat{V} \hat{W}} = \tilde{\omega}_{U
\hat{V} \hat{W}} \pm \frac{1}{2} H_{\hat{y} U \hat{V} \hat{W}}
\end{equation}
and $\tilde{\omega}_{U \hat{V} \hat{W}} = e^X \, \!_{\hat{W}} \left(
\tilde{\Gamma}_U \, \!^Y \, \!_X e_{Y \hat{V}} - \partial_U e_{X \hat{V}}
\right)$ is the Levi-Civita connection for the vielbein $\tilde{e}_{U
\hat{V}}$, that satisfies $\tilde{e}_{U \hat{W}} \tilde{e}_{V \hat{X}}
\eta^{\hat{W} \hat{X}} = \tilde{G}_{U V}$, where $\eta^{\hat{U} \hat{V}}$ is
the Minkowski metric on the $9 + 1$ dimensional boundary. \ The sign choice in
(\ref{omega bar}) is correlated with the chirality conditions on the
gravitino, gaugino, and supersymmetry variation parameter on the boundary.
Variation of the metric in $S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{HW}}}$ leads to the Einstein equations:
\begin{equation}
\label{Einstein equations} R_{I J} - \frac{1}{2} RG_{I J} - \frac{1}{12} H_I
\, \!^{K L M} H_{J K L M} + \frac{1}{96} H^{K L M N} H_{K L M N} G_{I J} = 0
\end{equation}
and on the boundary to the Israel boundary conditions {\cite{Israel, Chamblin
Reall, Dyer Hinterbichler, Moss 2}}:
\begin{equation}
\label{Israel boundary conditions} K^{U V} - K \tilde{G}^{U V} -
\kappa_{11}^2 \tilde{T}^{\left( \mathrm{{bos}} \right) U V} = 0,
\end{equation}
where
\begin{equation}
\label{T tilde i U V} \tilde{T}^{\left( \mathrm{{bos}} \right) U V} =
\frac{2}{\tilde{e}} \frac{\delta S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{YM}}}}{\delta \tilde{G}_{U V}} .
\end{equation}
The boundary is equivalent to a double-sided mirror at $y = y_1$, because all
the fields on one side of the mirror are exactly copied, up to sign, on the
other side of the mirror. \ The Yang-Mills multiplet is adjacent to the
mirror, but infinitesimally displaced from it, so that it has its own
reflection infinitesimally on the other side of the mirror {\cite{Lu}}. \ The
bulk and its reflection on the other side of the mirror form two parts of a
single closed manifold, and the formula (\ref{CJS action}) for the classical
action in the bulk applies on both sides of the mirror, with the \emph{same}
$\mathrm{SO}\left(10,1\right) $-invariant tensor $\epsilon^{\hat{I}_1 \ldots
\hat{I}_{11}}_{11}$ occurring in the definition of $\epsilon^{I_1 \ldots
I_{11}}_{11}$ on both sides of the mirror. Thus $\epsilon^{\hat{I}_1 \ldots
\hat{I}_{11}}_{11}$ transforms as a \emph{pseudo-tensor} under reflection in
the mirror, which means that its components are multiplied by an extra factor
of $-1$ under the reflection, relative to what they would have become if it
had transformed as a tensor, because it would have been multiplied by $-1$
under the reflection, if it had transformed as a tensor.
The metric $G_{I J}$ transforms as a tensor under the reflection, because if
it transformed as a pseudo-tensor, the measure factor $\sqrt{-G}$ would become
imaginary under the reflection, and the Ricci scalar would change sign under
the reflection. The Yang-Mills gauge field $A_U^{\mathcal{A}}$ also
transforms as a tensor under the reflection, because it only has components
tangential to the boundary, and if it transformed as a pseudo-tensor, its
components would have to vanish identically, if they were continuous across
the mirror.
However to preserve the sign of the last term, called the Chern-Simons term,
of the classical action in the bulk (\ref{CJS action}) under reflection in the
mirror, the 3-form $C_{I J K}$ has to transform, like $\epsilon^{\hat{I}_1
\ldots \hat{I}_{11}}_{11}$, as a pseudo-tensor under the reflection.
Thus $C_{U V W}$ and $H_{U V W X}$
change sign on reflection in the mirror, so if $H_{U V W X}$ is continuous
across the mirror, $H_{U V W X} = 0$ at the mirror. \ Thus the boundary
conditions for $C_{I J K}$ are controlled not by the action, but by the
pseudo-tensor transformation rule of $C_{I J K}$ under reflection in the
mirror, and continuity. \ Therefore variations of $C_{I
J K}$ in $S_{\mathrm{{HW}}}$ have $\delta C_{U V W} = 0$ at the boundary,
and lead only to the 3-form field equations, whose bosonic terms are:
\begin{equation}
\label{3 form field equations} D_L H^{L I J K} - \frac{1}{3456} \epsilon^{I
J K L M N O P Q R S}_{11} H_{L M N O} H_{P Q R S} = 0
\end{equation}
At zeroth order in $\kappa_{11}$, $S_{\mathrm{{YM}}}$ is absent and $H_{U
V W X}$ is continuous across the mirror, so the boundary condition is $H_{U V
W X} = 0$. \ At order $\kappa^{2 / 3}_{11}$, the $E_8$ Yang-Mills multiplet is
required to cancel the gravitational anomalies resulting from the chirality of
the gravitino boundary condition, and $H_{U V W X}$ is no longer continuous
across the mirror. \ The boundary condition for $H_{I J K L}$ at the surface
of the mirror is determined by anomaly cancellation and supersymmetry, and the
bosonic terms are:
\begin{equation}
\label{bc for H} H_{U V W X} = \pm \frac{3}{2 \pi} \left(
\frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left( \frac{1}{30}
\mathrm{{tr}} F_{\left[ U V \right.} F_{\left. W X \right]} -
\frac{1}{2} \bar{R}_{\left[ U V \right.} \, \!^{\hat{Y} \hat{Z}}
\bar{R}_{\left. W X \right] \hat{Y} \hat{Z}} - \frac{1}{3} \partial_{\left[
U \right.} \Omega_{\left. V W X \right]}^{\left( \mathrm{H} \right)} \right)
\end{equation}
where
\[ \Omega_{U V W}^{\left( \mathrm{H} \right)} = 3 H_{y \left[ U \right.} \,
\!^{X_1 X_2} D_V H_{\left. W \right] X_1 X_2 y} + 3 H_{y X_1 X_2 \left[ U
\right.} D^{X_1} H^{X_2} \, \!_{\left. V W \right] y} \]
\begin{equation}
\label{Omega H} \hspace{1cm} - \frac{1}{36} \varepsilon_{U V W} \, \!^{X_1
X_2 X_3 X_4 X_5 X_6 X_7} H_{y X_1 X_2 X_3} D_{X_4} H_{X_5 X_6 X_7 y}
\end{equation}
The sign choice in (\ref{bc for H}) is correlated with the chirality
conditions on the gravitino, gaugino, and supersymmetry variation parameter on
the boundary. \ Square brackets around two or more indices denote
antisymmetrization with unit total weight.
The boundary condition (\ref{bc for H}) can be integrated to:
\begin{equation}
\label{bc for C} C_{U V W} = \pm \frac{1}{4 \pi} \left( \frac{\kappa_{11}}{4
\pi} \right)^{2 / 3} \left( \frac{1}{30} \Omega_{U V W}^{\left( \mathrm{Y}
\right)} - \frac{1}{2} \Omega_{U V W}^{\left( \mathrm{L} \right)} -
\frac{1}{2} \Omega_{U V W}^{\left( \mathrm{H} \right)} \right) + \lambda_{U
V W},
\end{equation}
where
\begin{equation}
\label{Omega Y} \Omega_{U V W}^{\left( \mathrm{Y} \right)} = 6
\mathrm{{tr}} \left( A_{\left[ U \right.} \partial_V A_{\left. W
\right]} + \frac{2}{3} iA_{\left[ U \right.} A_V A_{\left. W \right]}
\right)
\end{equation}
and
\begin{equation}
\label{Omega L} \Omega_{U V W}^{\left( \mathrm{L} \right)} = - 6 \left(
\bar{\omega}_{\left[ U \right.} \, \!^{\hat{X}_1 \hat{X}_2} \partial_V
\bar{\omega}_{\left. W \right] \hat{X}_2 \hat{X}_1} + \frac{2}{3}
\mathcal{A}_{\left\{ U, V, W \right\}} \bar{\omega}_U \, \!^{\hat{X}_1} \,
\!_{\hat{X}_2} \bar{\omega}_V \, \!^{\hat{X}_2} \, \!_{\hat{X}_3}
\bar{\omega}_W \, \!^{\hat{X}_3} \, \!_{\hat{X}_1} \right)
\end{equation}
are Yang-Mills and Lorentz Chern-Simons 3-forms respectively, $\lambda_{U V
W}$ is an arbitrary closed 3-form on the boundary, and the notation
$\mathcal{A}_{\left\{ U, V, W \right\}}$ indicates that what follows it is to
be antisymmetrized in the indices $U, V, W$, with unit total weight.
The classical field equations of $d = 11$ supergravity do not have any
solutions with 4 flat extended dimensions if $\mathcal{M}^7$ has a compact
hyperbolic factor, for any arrangement of 4-form fluxes such that the flux
bilinears are on average covariantly constant, so it is necessary to include
the leading quantum corrections to the classical action in the bulk. \ The
quantum-corrected field equations have the form $\frac{\delta
\Gamma_{\mathrm{{HW}}}}{\delta \Phi} = 0$, where $\Phi$ represents all
the fields, and the bosonic part of the quantum effective action
$\Gamma_{\mathrm{{HW}}}$ {\cite{Schwinger, DeWitt, Nambu Jona Lasinio 1,
Nambu Jona Lasinio 2}} has the form:
\begin{equation}
\label{Gamma HW} \Gamma^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{HW}}} = S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{HW}}} + \Gamma^{\left( 8, \mathrm{{bos}}
\right)}_{\mathrm{{SG}}} + \Gamma^{\left( \mathrm{h.o., {bos}}
\right)}_{\mathrm{{HW}}} .
\end{equation}
Here $\Gamma^{\left( 8, \mathrm{{bos}} \right)}_{\mathrm{{SG}}}$ is
a dimension 8 local polynomial in the fields and derivatives in the bulk, of
order $\kappa^{4 / 3}_{11}$ relative to $S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{CJS}}}$, and $\Gamma^{\left( \mathrm{h.o., {bos}}
\right)}_{\mathrm{{HW}}}$, representing the higher order corrections,
contains both local and non-local terms, and is of order $\kappa^2_{11}$ or
higher relative to $S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{CJS}}}$ in the bulk, and of order $\kappa^{4 /
3}_{11}$ or higher relative to $S^{\left( \mathrm{{bos}}
\right)}_{\mathrm{{GH}}}$ on the boundary. \ The dimension of a local
monomial in the fields and derivatives is defined by counting 0 for each
metric or 3-form, $\frac{1}{2}$ for each gravitino, and 1 for each derivative,
and on the boundary also counting 1 for each Yang-Mills gauge field and
$\frac{3}{2}$ for each gaugino.
The leading quantum corrections to $d = 11$ supergravity in the bulk have the
form {\cite{Tseytlin}}:
\begin{equation}
\label{Gamma 8 SG} \Gamma^{\left( 8, \mathrm{{bos}}
\right)}_{\mathrm{{SG}}} = \frac{1}{147456 \pi^2 \kappa_{11}^2} \left(
\frac{\kappa_{11}}{4 \pi} \right)^{4 / 3} \int_{\mathcal{B}} d^{11} xe
\left( t_8 t_8 R^4 - \frac{1}{4!} \epsilon_{11} \epsilon_{11} R^4 -
\frac{1}{6} \epsilon_{11} t_8 CR^4 + \Xi^{\left( \mathrm{{flux}}
\right)} \right)
\end{equation}
Here for an arbitrary tensor $X_{I J K L}$, that is antisymmetric in its first
two indices and antisymmetric in its last two indices:
\begin{equation}
\label{t8 t8 R4} t_8 t_8 X^4 \equiv t^{I_1 I_2 J_1 J_2 K_1 K_2 L_1 L_2}_8
t^{M_1 M_2 N_1 N_2 O_1 O_1 P_1 P_2}_8 X_{I_1 I_2 M_1 M_2} X_{J_1 J_2 N_1
N_2} X_{K_1 K_2 O_1 O_2} X_{L_1 L_2 P_1 P_2},
\end{equation}
where $t^{I J K L M N O P}_8$ is a tensor built from $G^{I J}$ and
antisymmetric in each successive pair of indices, such that for antisymmetric
tensors $A_{I J}$, $B_{I J}$, $C_{I J}$, and $D_{I J}$:
\[ t^{I J K L M N O P}_8 A_{I J} B_{K L} C_{M N} D_{O P} = \hspace{6cm} \]
\[ = 8 \left( \mathrm{{tr}} \left( ABCD \right) + \mathrm{{tr}}
\left( ACBD \right) + \mathrm{{tr}} \left( ACDB \right) \right) \]
\[ - 2 \left( \mathrm{{tr}} \left( AB \right) \mathrm{{tr}} \left(
CD \right) + \mathrm{{tr}} \left( AC \right) \mathrm{{tr}} \left(
BD \right) + \mathrm{{tr}} \left( AD \right) \mathrm{{tr}} \left(
BC \right) \right) = \]
\[ = 8 \left( A_{I J} B_{J K} C_{K L} D_{L I} + A_{I J} C_{J K} B_{K L} D_{L
I} + A_{I J} C_{J K} D_{K L} B_{L I} \right) \]
\begin{equation}
\label{t8 A B C D} - 2 \left( A_{I J} B_{J I} C_{K L} D_{L K} + A_{I J} C_{J
I} B_{K L} D_{L K} + A_{I J} D_{J I} B_{K L} C_{L K} \right)
\end{equation}
Here and in the following, repeated lower coordinate indices are understood to
be contracted with an inverse metric, for example $A_{I J} B_{J K} \equiv A_I
\, \!^J B_{J K} = G^{J L} A_{I L} B_{J K}$. \ Thus for an arbitrary such
tensor $X_{I J K L}$:
\begin{dmath}[compact, spread=3pt]
\label{t8 t8 X4} t_8 t_8 X^4 = 12 \hspace{0.25em} X_{I J K L} X_{J I L K}
X_{M N O P} X_{N M P O} + 24 \hspace{0.25em} X_{I J K L} X_{J I M N} X_{O P
L K} X_{P O N M} - 96 \hspace{0.25em} X_{I J K L} X_{J I L M} X_{N O M P}
X_{O N P K} - 48 \hspace{0.25em} X_{I J K L} X_{J I M N} X_{O P L M} X_{P O
N K} - 96 \hspace{0.25em} X_{I J K L} X_{J M L K} X_{M N O P} X_{N I P O} -
48 \hspace{0.25em} X_{I J K L} X_{J M N O} X_{M P L K} X_{P I O N} + 192
\hspace{0.25em} X_{I J K L} X_{J M L N} X_{M O N P} X_{O I P K} + 384
\hspace{0.25em} X_{I J K L} X_{J M L N} X_{M O P K} X_{O I N P}
\end{dmath}
And similarly, for an arbitrary such tensor $X_{I J K L}$:
\begin{dmath}[compact, spread=3pt]
\label{eps11 eps11 X4} \frac{1}{4!} \epsilon_{11} \epsilon_{11} X^4 \equiv
\frac{1}{4!} \epsilon_{11}^{I J K L_1 L_2 M_1 M_2 N_1 N_2 O_1 O_2} \left(
\epsilon_{11} \right)_{I J K P_1 P_2 Q_1 Q_2 R_1 R_2 S_1 S_2} \times
X_{L_1 L_2} \, \!^{P_1 P_2} X_{M_1 M_2} \, \!^{Q_1 Q_2} X_{N_1 N_2} \,
\!^{R_1 R_2} X_{O_1 O_2} \, \!^{S_1 S_2} = - \frac{8!}{4} X_{L_1 L_2} \,
\!^{\left[ L_1 L_2 \right.} X_{M_1 M_2} \, \!^{M_1 M_2} X_{N_1 N_2} \,
\!^{N_1 N_2} X_{O_1 O_2} \, \!^{\left. O_1 O_2 \right]}
\end{dmath}
And similarly:
\begin{dmath}
\label{eps11 t8 C R4} \frac{1}{6} \epsilon_{11} t_8 CR^4 \equiv 4
\epsilon^{I_1 \ldots I_{11}} C_{I_1 I_2 I_3} R_{I_4 I_5 J K} R_{I_6 I_7 K L}
R_{I_8 I_9 L M} R_{I_{10} I_{11} M J} - \epsilon^{I_1 \ldots I_{11}} C_{I_1
I_2 I_3} R_{I_4 I_5 J K} R_{I_6 I_7 K J} R_{I_8 I_9 L M} R_{I_{10} I_{11} M
L}
\end{dmath}
The relative coefficients of all terms in (\ref{Gamma 8 SG}) are fixed by
supersymmetry, up to the fact that arbitrary multiples of linear combinations
of terms that vanish when the classical field equations (\ref{Einstein
equations}) and (\ref{3 form field equations}) are satisfied can be added,
because the overall coefficients of such linear combinations of terms can be
adjusted arbitrarily by small amounts, by making redefinitions of the fields
of the form $G_{I J}, C_{I J K} \rightarrow G_{I J} + X_{I J}, C_{I J K} +
Y_{I J K}$, where $X_{I J}$ and $Y_{I J K}$ are polynomials in the fields and
their derivatives, with small coefficients {\cite{Hyakutake Ogushi 2, Hyakutake,
Metsaev}}. \ The coefficient of the $\epsilon_{11} t_8 CR^4$ term, which is
known as the Green-Schwarz term because of its role in anomaly cancellation
{\cite{Green Schwarz}}, is fixed absolutely by anomaly cancellation on
five-branes {\cite{Duff Liu Minasian, Witten 5 branes, Freed Harvey Minasian
Moore, Bilal Metzger, Harvey}}, and confirmed by anomaly cancellation in
Ho\v{r}ava-Witten theory {\cite{de Alwis, Conrad, Faux Lust Ovrut, Lu, Bilal
Derendinger Sauser, Harmark, Bilal Metzger, Meissner Olechowski, Moss 3}}.
The flux term $\Xi^{\mathrm{{flux}}}$ in (\ref{Gamma 8 SG}) is a linear
combination of dimension 8 coordinate scalar monomials built from $R_{I J K
L}$, $G^{I J}$, $\epsilon^{I_1 \ldots I_{11}}_{11}$, $H_{I J K L}$, and the
covariant derivative $D_I$, such that $H_{I J K L}$ occurs at least once in
each monomial, and the number of $R_{I J K L}$ plus the number of $H_{I J K
L}$ is at least 4. \ $\epsilon^{I_1 \ldots I_{11}}_{11}$ occurs once in each
monomial with an odd number of $D_I$, and is absent from monomials with an
even number of $D_I$. \ The only building blocks of odd dimension are $H_{I J
K L}$ and $D_I$, so each occurs an odd number of times if the other does. \ If
$H_{I J K L}$ occurs exactly once then there are three $R_{I J K L}$'s and one
$D_I$, hence also an $\epsilon^{I_1 \ldots I_{11}}_{11}$, and it is impossible
to build a coordinate scalar that does not vanish identically by a Bianchi
identity or the Ricci cyclic identity {\cite{Deser Seminara 1, Deser Seminara
2}}, so in fact there are at
least two $H_{I J K L}$'s in each monomial.
The terms in $\Xi^{\mathrm{{flux}}}$ can be classified by the numbers of
$R_{I J K L}$ and $H_{I J K L}$ in them. \ For general fluxes $H_{I J K L}$
only the $R^2 H^2$, $RH^3$, and $H^4$ terms are known at present {\cite{Deser
Seminara 1, Deser Seminara 2, Peeters Plefka Stern, Metsaev}}. \ These terms
simplify substantially in the special case where there is a particular
coordinate, which for convenience of notation I shall call $y$, although in
general there is no HW boundary nearby, such that all nonvanishing components
of $H_{I J K L}$ have an index $y$, all the fields are independent of $y$,
and, borrowing the notation used above for HW theory, the metric components
$G_{U y}$ are 0. \ Then the $RH^3$ terms vanish, because they have two $H_{I J
K L}$'s contracted with an $\epsilon^{I_1 \ldots I_{11}}_{11}$ {\cite{Deser
Seminara 1, Deser Seminara 2}}, and by the correspondence with type IIA
superstring theory {\cite{Green Schwarz IIA, Witten Various Dimensions}}, in
the special case where only the fields of the type I supergravity multiplet in
10 dimensions are nonzero, the 4-field terms in $\Gamma^{\left( 8,
\mathrm{{bos}} \right)}_{\mathrm{{SG}}}$ must reduce to the 4-field
terms in:
\begin{dmath}
\label{Gamma I y} \left. \Gamma^{\left( 8, \mathrm{{bos}}
\right)}_{\mathrm{{SG}}} \right|_{\mathrm{I}, y} = \frac{1}{147456
\pi^2 \kappa_{11}^2} \left( \frac{\kappa_{11}}{4 \pi} \right)^{4 / 3}
\int_{\mathcal{B}} d^{11} xe \left( t_8 t_8 \bar{R}^4 - \frac{1}{4!}
\epsilon_{11} \epsilon_{11} \bar{R}^4 \right),
\end{dmath}
where $\bar{R}_{U V W X} = e_W \, \!^{\hat{Y}} e_X \, \!^{\hat{Z}} \bar{R}_{U
V \hat{Y} \hat{Z}}$, with $\bar{R}_{U V \hat{W} \hat{X}}$ defined by (\ref{R
bar}) and $\bar{\omega}_{U \hat{V} \hat{W}}$ defined by (\ref{omega bar}),
$\bar{R}_{y U V W} = \bar{R}_{V W y U} = R_{y U V W} = 0$, $\bar{R}_{y U y V}
= R_{y U y V}$, and either sign in (\ref{omega bar}) may be used, since with
$\tilde{H}_{U V W} \equiv H_{\hat{y} U V W}$, the $\tilde{R}_{U V W X}$ and\\
$-
\frac{1}{4} \tilde{G}^{Y Z} \left( \tilde{H}_{U W Y} \tilde{H}_{V X Z} -
\tilde{H}_{V W Y} \tilde{H}_{U X Z} \right)$ terms in $\bar{R}_{U V W X}$ are
unaltered by swapping $U V$ with $W X$, while the $\pm \frac{1}{2} \left(
\tilde{D}_U \tilde{H}_{V W X} - \tilde{D}_V \tilde{H}_{U W X} \right)$ terms
change sign due to the Bianchi identity for $H_{I J K L}$, so reversing the
sign of $\tilde{H}_{U V W}$ is equivalent to swapping the two $t_8$'s or the
two $\epsilon_{11}$'s in (\ref{Gamma I y}), so (\ref{Gamma I y}) only contains
even powers of $\tilde{H}_{U V W}$. \ The coefficient of $\tilde{H}_{U V W}$
in $\bar{\omega}_{U \hat{V} \hat{W}}$ is fixed to $\pm \frac{1}{2}$, in
agreement with (\ref{omega bar}), by comparison with equations (2.8), (2.12),
and (2.13) of {\cite{Gross Sloan}}, noting the unusual normalization of
antisymmetrization brackets around indices used in that article, as shown by a
statement after their equation (2.9), or equivalently by comparison with
equations (5.10), (5.18), and (5.36) of {\cite{Policastro Tsimpis}}, who use
the standard normalization of antisymmetrization brackets, as defined after
equation (\ref{Omega H}) above.
Richards {\cite{Richards 2}} has found evidence that for the field
configurations specified above, the formula (\ref{Gamma I y}) for $\left.
\Gamma^{\left( 8, \mathrm{{bos}} \right)}_{\mathrm{{SG}}}
\right|_{\mathrm{I}, y}$ might also be valid beyond the 4 field level, in
agreement with an earlier suggestion by Kehagias and Partouche {\cite{KP}}.
However the following example shows that (\ref{Gamma I y}) as it stands cannot
be oxidized to a generally covariant formula in 11 dimensions.
We consider the Cartesian product of 7-dimensional Minkowski space and a
2-sphere of radius $a$ and a 2-sphere of radius $b$. \ We choose spherical
polar coordinates $\left( \theta, \varphi \right)$ on the first 2-sphere, and
spherical polar coordinates $\left( \eta, \xi \right)$ on the second 2-sphere,
so the metric is:
\begin{equation}
\label{metric for KP example} ds_{11}^2 = G_{IJ} dx^I dx^J = \eta_{\alpha
\beta} dx^{\alpha} dx^{\beta} + a^2 d \theta^2 + a^2 \mathrm{\sin}^2 \theta
d \varphi^2 + b^2 d \eta^2 + b^2 \mathrm{\sin}^2 \eta d \xi^2,
\end{equation}
where $\eta_{\alpha \beta} = \mathrm{{diag}} \left( - 1, 1, \ldots, 1
\right)$ is the metric on $\left( 6 + 1 \right)$-dimensional Minkowski space.
\ The non-vanishing components of $H_{I J K L}$, up to index permutations,
are:
\begin{equation}
\label{flux for KP example} H_{\theta \varphi \eta \xi} = h\,
\varepsilon_{\theta \varphi}\, \varepsilon_{\eta \xi} =h\, \mathrm{\sin} \theta
\, \mathrm{\sin} \eta,
\end{equation}
where if indices $i, j, \ldots$ are tangential to the first 2-sphere and
indices $p, q, \ldots$ are tangential to the second 2-sphere, $\varepsilon_{i
j}$ is a covariantly constant harmonic 2-form on the first 2-sphere, with
non-vanishing components $\varepsilon_{\theta \varphi} = -
\varepsilon_{\varphi \theta} = \mathrm{\sin} \theta$, and $\varepsilon_{p q}$
is a covariantly constant harmonic 2-form on the second 2-sphere, with
non-vanishing components $\varepsilon_{\eta \xi} = - \varepsilon_{\xi \eta} =
\mathrm{\sin} \eta$. \ The flux (\ref{flux for KP example}) is a covariantly
constant harmonic 4-form on the Cartesian product of the two 2-spheres.
For the metric (\ref{metric for KP example}) and the flux (\ref{flux for KP
example}), either $\varphi$ or $\xi$ could be the special coordinate $y$ in
(\ref{Gamma I y}). The $\epsilon_{11} \epsilon_{11} \bar{R}^4$ term gives 0
because $\bar{R}_{I J K L}$ is 0 unless all its indices are tangential to the
Cartesian product of the two 2-spheres, and if we choose $\varphi$ as $y$ then
using Cadabra {\cite{Cadabra 1, Cadabra 2, Cadabra 3, Cadabra 4, Cadabra 5,
Cadabra 6}} we find:
\begin{equation}
\label{t8 t8 for KP example} t_8 t_8 \bar{R}^4 = \frac{576}{a^8} +
\frac{384}{a^4 b^4} + \frac{576}{b^8} - \frac{192 h^2}{a^8 b^6} - \frac{576
h^2}{a^4 b^{10}} + \frac{72 h^4}{a^{12} b^8} + \frac{264 h^4}{a^8 b^{12}} -
\frac{60 h^6}{a^{12} b^{14}} + \frac{45 h^8}{4 a^{16} b^{16}}
\end{equation}
Choosing $\xi$ as $y$ gives (\ref{t8 t8 for KP example}) with $a$ and $b$
swapped, so since (\ref{t8 t8 for KP example}) is not symmetrical in $a$ and
$b$, (\ref{Gamma I y}) cannot be oxidized to a generally covariant expression
in 11 dimensions, because any such expression would give a result symmetrical
in $a$ and $b$.
Thus (\ref{Gamma I y}) must be subject to a correction if used beyond the
4-field level. \ I shall consider only the case where $H_{I J K L}$ is
covariantly constant. \ For a metric $G_{I J}$ as specified before (\ref{Gamma
I y}), the non-vanishing Christoffel symbols are:
\[ \Gamma_U \, \!^V \, \!_W = \frac{1}{2} G^{V X} \left( \partial_U G_{X W} +
\partial_W G_{X U} - \partial_X G_{U W} \right) = \tilde{\Gamma}_U \, \!^V
\, \!_W \]
\begin{equation}
\label{Christoffel symbols for KP case} \Gamma_y \, \!^y \, \!_W =
\frac{1}{2} G^{y y} \partial_W G_{y y}, \hspace{1cm} \Gamma_y \, \!^V \,
\!_y = - \frac{1}{2} G^{V X} \partial_X G_{y y}, \hspace{1cm} \Gamma_U \,
\!^y \, \!_y = \frac{1}{2} G^{y y} \partial_U G_{y y}
\end{equation}
Thus with $H_{\hat{y} V W X} = \frac{1}{\sqrt{G_{y y}}} H_{y V W X}$:
\begin{equation}
\label{covariant constancy in 10 dims and 11 dims} \tilde{D}_U H_{\hat{y} V
W X} = \frac{1}{\sqrt{G_{y y}}} D_U H_{y V W X}
\end{equation}
So $H_{\hat{y} V W X}$ is covariantly constant as a 3-form in the
10-dimensional sense, with the $\hat{y}$ index ignored, if and only if $H_{y V
W X}$ is covariantly constant as a 4-form in the 11-dimensional sense.
If the 4-form is covariantly constant, then from (\ref{R bar}) and (\ref{omega
bar}):
\begin{equation}
\label{R bar for covariantly constant H} \bar{R}_{U V W X} = R_{U V W X} -
\frac{1}{4} G^{y y} G^{Y Z} H_{y Y U W} H_{y Z V X} + \frac{1}{4} G^{y y}
G^{Y Z} H_{y Y V W} H_{y Z U X}
\end{equation}
Thus a simple guess for the oxidation of (\ref{Gamma I y}) to 11 dimensions
when the 4-form is covariantly constant would be to replace $R_{I J K L}$ in
(\ref{Gamma 8 SG}) by:
\begin{equation}
\label{X I J K L} X_{I J K L} = R_{I J K L} - \frac{1}{8} H_{I K M N} H_{J
L} \, \!^{M N} + \frac{1}{8} H_{J K M N} H_{I L} \, \!^{M N},
\end{equation}
and discard $\Xi^{\left( \mathrm{{flux}} \right)}$, because terms with a
covariant derivative acting on the 4-form now vanish, and terms without
covariant derivatives are absorbed in the $X^4$ terms. \ This gives a
correction to (\ref{Gamma I y}) when only the fields of the type I
supergravity multiplet in 10 dimensions are nonzero, as required by the above
example, because we then have:
\begin{equation}
\label{X y U y V} X_{y U y V} = R_{y U y V} + \frac{1}{8} H_{U y M N} H_{y
V} \, \!^{M N}
\end{equation}
The correction term $ \frac{1}{8} H_{U y M N} H_{y V} \, \!^{M N} $ occurs
multiplied by $2 G^{y y}$ as a source term in the
classical Einstein equations (\ref{Einstein equations}) for $G_{U V}$, so
terms containing it might arise from a modified treatment of the
classical Einstein equation in extracting the $d = 10$ effective action from
the superstring scattering amplitude.
\section{The bulk}
\label{The bulk}
I shall now look for solutions of the quantum-corrected $d = 11$ Einstein
equations that have 4 flat extended dimensions and a compact 7-manifold
$\mathcal{M}^7$ with one or more compact hyperbolic factors, in the presence
of magnetic 4-form fluxes whose bilinears are on average covariantly constant.
\ The averages of terms containing covariant derivatives of the 4-form could
in principle be non-zero, but I shall assume that these terms can be neglected
for a first approximation. \ As a guess for the approximate oxidation of the
Kehagias-Partouche form (\ref{Gamma I y}) when covariant derivatives of the
4-form can be neglected, I shall use (\ref{Gamma 8 SG}) with $R_{I J K L}$
replaced by the $X_{I J K L}$ defined in (\ref{X I J K L}), and $\Xi^{\left(
\mathrm{{flux}} \right)}$ discarded. \ I shall consider $\mathcal{M}^7$
that are Cartesian products of compact hyperbolic and spherical factors, and
work in the leading order of the Lukas-Ovrut-Waldram harmonic expansion of the
energy-momentum tensor on $\mathcal{M}^7$ {\cite{Lukas Ovrut Waldram}}.
There is then one independent Einstein equation for each Cartesian factor of
$\mathcal{M}^7$, plus one Einstein equation for the 4 flat extended
dimensions, because no source terms for the off-diagonal blocks of the Ricci
tensor can be built from the covariantly constant average flux bilinears, and
the source term for each diagonal block of the Ricci tensor is proportional to
the corresponding diagonal block of the metric. \ The Einstein equations are
consistent with the symmetry of the metric ansatz, so by Palais's Principle of
Symmetric Criticality {\cite{Palais, Deser Franklin, Torre}}, the Einstein
equations can be derived by varying the action with respect to an independent
scale factor introduced for each Cartesian factor of $\mathcal{M}^7$, and a
scale factor for the 4 flat extended dimensions.
Within these approximations we can also introduce 5-branes {\cite{Guven}},
which are magnetic sources such that the 5-form dual to the 6-dimensional
world-volume of the 5-brane provides a source for the Bianchi identity of the
4-form, localized on the 5-brane world-volume. \ To preserve Lorentz
invariance 4 dimensions of the 5-brane world-volume must lie along the 4
extended dimensions, and the remaining two dimensions of the 5-brane
world-volume wrap a 2-cycle of $\mathcal{M}^7$. \ I do not consider 2-branes
{\cite{Duff Stelle}} since they are electric sources and their presence would
violate Lorentz invariance, and I do not consider a flux $H_{\mu \nu \sigma
\tau}$ proportional to $\epsilon_{\mu \nu \sigma \tau}$ with indices along the
4 extended dimensions, because by the HW boundary condition (\ref{bc for H})
it would require the presence either of Lorentz-violating Yang-Mills fields,
or Lorentz-violating fluxes, or fluxes that violate translation-invariance
along the extended dimensions, and it is not in general simultaneously
measurable with the magnetic fluxes \cite{Moore}.
The simplest case is when $H^7$ is a compact hyperbolic 7-manifold
$\bar{H}^7$. \ The metric ansatz (\ref{metric ansatz}) now takes the form:
\begin{equation}
\label{metric ansatz for H7} ds_{11}^2 = G_{IJ} dx^I dx^J = A^2 \eta_{\mu
\nu} dx^{\mu} dx^{\nu} + B^2 g_{i j} dx^i dx^j,
\end{equation}
where indices $i, j, \ldots$ are tangential to the $\bar{H}^7$, and $g_{i j}$
is a metric of sectional curvature $- 1$ on the $\bar{H}^7$. \ We introduce
fluxes $H_{i j k l}$ wrapping 4-cycles of $\bar{H}^7$ such that on average:
\begin{equation}
\label{flux bilinears} H_{i j m n} H_{k l o p} G^{m o} G^{n p} =
\frac{h^2}{B^8} \left( G_{i k} G_{j l} - G_{j k} G_{i l} \right),
\end{equation}
where $h$ is a constant. \ The dependence on $B$ is fixed by the fact that
$H_{i j m n}$ is independent of $B$, because the integral of $H_{i j k l} dx^i
dx^j dx^k dx^l$ over a 4-cycle of $\bar{H}^7$ is quantized, and independent of
$B$. \ From (\ref{flux bilinears}) and (\ref{X I J K L}) we find:
\begin{equation}
\label{X i j k l} X_{i j k l} = \left( G_{i l} G_{j k} - G_{j l} G_{i k}
\right) \left( \frac{1}{B^2} + \frac{h^2}{8 B^8} \right)
\end{equation}
The $\epsilon_{11} \epsilon_{11} X^4$ term (\ref{eps11 eps11 X4}) in the
action vanishes because the indices $i, j, \ldots$ run over only 7 dimensions,
and using Cadabra {\cite{Cadabra 1, Cadabra 2, Cadabra 3, Cadabra 4, Cadabra 5,
Cadabra 6}} to evaluate the
$t_8 t_8 X^4$ term (\ref{t8 t8 X4}), the action density is:
\begin{equation}
\label{action density for H7} \frac{A^4 B^7}{2 \kappa^2_{11}} \left( - 42
\frac{1}{B^2} - \frac{42}{48} \frac{h^2}{B^8} - \frac{k^2}{B^5} +
\frac{254016}{73728 \pi^2} \left( \frac{\kappa_{11}}{4 \pi} \right)^{4 / 3}
\left( \frac{1}{B^2} + \frac{h^2}{8 B^8} \right)^4 \right)
\end{equation}
The $- \frac{k^2}{B^5}$ term is the classical action of a distribution of
5-branes. \ There is evidence that in HW theory, a 5-brane can only have the
fundamental tension $T_5 = \frac{1}{8 \pi} \left( \frac{4 \pi}{\kappa_{11}}
\right)^{4 / 3}$ {\cite{Lu, Harmark}}. \ The 5-brane extends with any cycle
that it wraps, and the denominator factor is to cancel the measure factor for
each cycle it does not wrap.
To analyse the field equations it is convenient to introduce rescaled
parameters
\begin{equation}
\label{rescaled parameters} \tilde{B} = \frac{2^{\frac{29}{18}}
\pi^{\frac{5}{9}}}{21^{\frac{1}{3}}} \, \frac{B}{\kappa_{11}^{2 / 9}},
\hspace{1.5cm} \tilde{h} = \frac{2^{\frac{29}{6}} \pi^{\frac{5}{3}}}{21} \,
\frac{h}{\kappa_{11}^{2 / 3}}, \hspace{1.5cm} \tilde{k} =
\frac{2^{\frac{29}{12}} \pi^{\frac{5}{6}}}{21^{\frac{1}{2}}} \,
\frac{k}{\kappa_{11}^{1 / 3}},
\end{equation}
in terms of
which the action density becomes a constant multiple of:
\begin{equation}
\label{with rescaled parameters} A^4 \tilde{B}^7 \left( - 42
\frac{1}{\tilde{B}^2} - \frac{42}{48} \frac{\tilde{h}^2}{\tilde{B}^8} -
\frac{\tilde{k}^2}{\tilde{B}^5} + \left( \frac{1}{\tilde{B}^2} +
\frac{\tilde{h}^2}{8 \tilde{B}^8} \right)^4 \right)
\end{equation}
The field equations are then:
\begin{equation}
4096 \tilde{B}^{27} \tilde{k}^2 - \tilde{h}^8 - 32 \tilde{B}^6
\tilde{h}^6 - 384 \tilde{B}^{12} \tilde{h}^4 + 3584 \tilde{B}^{24}
\tilde{h}^2 - 2048 \tilde{B}^{18} \tilde{h}^2 + 172032 \tilde{B}^{30} -
4096 \tilde{B}^{24} = 0
\end{equation}
\begin{dmath}[compact, spread=3pt]
8192 \tilde{B}^{27} \tilde{k}^2 + 25 \tilde{h}^8 + 608 \tilde{B}^6
\tilde{h}^6 + 4992 \tilde{B}^{12} \tilde{h}^4 - 3584 \tilde{B}^{24}
\tilde{h}^2 + 14336 \tilde{B}^{18} \tilde{h}^2 + 860160 \tilde{B}^{30} +
4096 \tilde{B}^{24} = 0
\end{dmath}
Adding these two equations, we find:
\begin{equation}
24 \left( 512 \tilde{B}^{27} \tilde{k}^2 + \tilde{h}^8 + 24 \tilde{B}^6
\tilde{h}^6 + 192 \tilde{B}^{12} \tilde{h}^4 + 512 \tilde{B}^{18}
\tilde{h}^2 + 43008 \tilde{B}^{30} \right) = 0
\end{equation}
Thus there is no solution. \ Consideration of the $\tilde{B}^{24} \tilde{h}^2$
and $\tilde{B}^{24}$ terms shows that this is the only linear combination, up
to an overall factor, that gives a non-negative coefficient to every term.
I then studied the $\bar{H}^5 \times \bar{H}^2$ and $\bar{H}^4 \times
\bar{H}^3$ cases, including all possible types of flux. \ In both cases, there
was a unique linear combination of the 3 field equations, up to an overall
factor, that gave a non-negative coefficient to every term, and thus proved
there was no solution. \ I then studied the $\bar{H}^5 \times S^2$, $\bar{H}^4
\times S^3$, and $\bar{H}^3 \times S^4$ cases, and also the $\bar{H}^3 \times
S^2 \times S^2$ case, with the radii of the two $S^2$'s set equal after
deriving the field equations. \ There was no longer any simple proof that no
solution existed, but searches by various methods found no solution. \ TeXmacs
files containing some of these calculations, together with Cadabra scripts and
source code for a C++ program used in one of the searches, are available on
the web page {\cite{h7bulk}}.
I therefore returned to the $\bar{H}^7$ case, and studied whether a field
redefinition could produce a solution, even though, in principle, field
redefinitions should have no physical effects when all orders of perturbation
theory are included. \ Trying a redefinition that affects only the coefficient
of the highest power of $\frac{1}{\tilde{B}}$ in the action, one finds that as
soon as this coefficient becomes negative, a solution appears. \ One might
suspect that a solution that disappears at the leading order it exists on
performing a field redefinition is unphysical, and that this will show up
through the presence of tachyonic modes or through large corrections at higher
orders in $\kappa^{2 / 3}_{11}$.
However there is no absolutely preferred set of coordinates in ``field
space'', and I will now show that a solution also appears when we make a field
redefinition that has the same form as the redefinition that transforms the
action with the unexpanded $\epsilon_{11} \epsilon_{11} X^4$ term to the form
that occurs naturally when supersymmetry is systematically implemented by the
Noether method {\cite{Hyakutake Ogushi 1, Hyakutake Ogushi 2, Hyakutake,
Hyakutake 2}}, except that instead of
making the redefinition with coefficient 1, we make it with a coefficient $< -
0.34$. \ Thus a solution appears after a redefinition that is much closer to
the identity than a redefinition that is certainly admissible.
To transform the action to the form that would naturally be obtained by the
Noether method, we have to remove all the Ricci terms from the expansion of
$\epsilon_{11} \epsilon_{11} X^4$ by using the classical Einstein equations
(\ref{Einstein equations}). \ Let:
\begin{equation}
\label{V I K} V_{I K} \equiv X_{I J K} \, \!^J = R_{I K} - \frac{1}{8} H_I
\, \!^{J M N} H_{K J M N}
\end{equation}
\begin{equation}
\label{U} U \equiv V_I \, \!^I = R - \frac{1}{8} H^{I J M N} H_{I J M N}
\end{equation}
where I used (\ref{X I J K L}). \ Let $\tilde{V}_{I K}$ and $\tilde{U}$ be the
result of using the classical Einstein equations (\ref{Einstein equations}) to
replace the Ricci terms in $V_{I K}$ and $U$ by $H^2$ terms. \ Then:
\begin{equation}
\label{V tilde I K} \tilde{V}_{I K} = - \frac{1}{24} H_I \, \!^{J M N} H_{K
J M N} - \frac{1}{144} H^{O L M N} H_{O L M N} G_{I K}
\end{equation}
\begin{equation}
\label{U tilde} \tilde{U} = - \frac{17}{144} H^{I J M N} H_{I J M N}
\end{equation}
So for the $\bar{H}^7$ case we find from (\ref{flux bilinears}):
\begin{equation}
\label{V tilde for H bar 7} \tilde{V}_{\mu \nu} = - \frac{7}{24}
\frac{h^2}{B^8} G_{\mu \nu}, \hspace{2cm} \tilde{V}_{i j} = - \frac{13}{24}
\frac{h^2}{B^8} G_{i j}
\end{equation}
\begin{equation}
\label{U tilde for H bar 7} \tilde{U} = - \frac{119}{24} \frac{h^2}{B^8}
\end{equation}
Now using Cadabra we find that:
\begin{dmath}[compact, spread=3pt]
\label{minus eps eps X4} - \frac{1}{4!} \epsilon_{11} \epsilon_{11} X^4 = 4
U^4 - 96 V_{I J} V_{I J} U^2 + 24 X_{I J K L} X_{I J K L} U^2 + 384 X_{I J K
L} V_{I K} V_{J L} U + 256 V_{I J} V_{I K} V_{J K} U - 384 X_{I J K L} X_{I
J K M} V_{L M} U + 32 X_{I J K L} X_{I J M N} X_{K L M N} U - 128 X_{I J K
L} X_{I M K N} X_{J N L M} U + 192 V_{I J} V_{I J} V_{K L} V_{K L} - 96 X_{I
J K L} X_{I J K L} V_{M N} V_{M N} - 768 X_{I J K L} X_{I M K N} V_{J L}
V_{M N} - 1536 X_{I J K L} V_{I K} V_{J M} V_{L M} + 768 X_{I J K L} X_{I J
K M} X_{L N M O} V_{N O} - 384 V_{I J} V_{I K} V_{J L} V_{K L} + 768 X_{I J
K L} X_{I J K M} V_{L N} V_{M N} + 384 X_{I J K L} X_{I J M N} V_{K M} V_{L
N} - 384 X_{I J K L} X_{I J M N} X_{K L M O} V_{N O} + 768 X_{I J K L} X_{I
M K N} V_{J M} V_{L N} + 1536 X_{I J K L} X_{I M K N} X_{J N L O} V_{M O} +
12 X_{I J K L} X_{I J K L} X_{M N O P} X_{M N O P} - 192 X_{I J K L} X_{I J
K M} X_{L N O P} X_{M N O P} + 24 X_{I J K L} X_{I J M N} X_{K L O P} X_{M N
O P} - 384 X_{I J K L} X_{I J M N} X_{K O M P} X_{L P N O} + 192 X_{I J K L}
X_{I M K N} X_{J O L P} X_{M O N P} - 384 X_{I J K L} X_{I M K N} X_{J O M
P} X_{L O N P}
\end{dmath}
This used the fact that $X_{I J K L}$ in (\ref{X I J K L}) has the same
monoterm symmetries as the Riemann tensor, namely antisymmetry in the first
two indices, antisymmetry in the last two indices, and symmetry under exchange
of the first two indices with the last two indices, although it does not
satisfy the Ricci cyclic identity, because it has a piece with [1111] Young
tableau symmetry.
Let $Z \equiv - \frac{1}{4!} \epsilon_{11} \epsilon_{11} X^4$, and let
$\tilde{Z}$ be the result of replacing $V_{I J}$ by $\tilde{V}_{I J}$ and $U$
by $\tilde{U}$ in the right-hand side of (\ref{minus eps eps X4}), so that $Z
= \tilde{Z}$ if the classical Einstein equations (\ref{Einstein equations})
are satisfied. \ Then $\left( \tilde{Z} - Z \right)$, with an arbitrary
coefficient that is not too large, is an expression that can be added to the
action at this order by means of a field redefinition. \ To reach the form of
the action that would naturally be obtained by imposing supersymmetry by the
Noether method, one has to add $\left( \tilde{Z} - Z \right)$ with coefficient
1 to the integrand of the dimension 8 local term (\ref{Gamma 8 SG}), on page
\pageref{Gamma 8 SG},
which is here being taken with $\Xi^{\left( \mathrm{{flux}} \right)}$
discarded, and $R_{IJKL}$ replaced by the $X_{I J K L}$ defined in (\ref{X I J
K L}), on page \pageref{X I J K L}.
From (\ref{V tilde for H bar 7}), (\ref{U tilde for H bar 7}), and (\ref{minus
eps eps X4}), we find using Cadabra and Maxima {\cite{Maxima}} that for the
$\bar{H}^7$ case, for which $Z = 0$:
\begin{equation}
\label{Z tilde minus Z} \tilde{Z} - Z = 7 \left( - \frac{32832}{B^8} +
\frac{17088 h^2}{B^{14}} + \frac{161 h^4}{2 B^{20}} + \frac{20395 h^6}{24
B^{26}} + \frac{2225255 h^8}{82944 B^{32}} \right)
\end{equation}
Thus adding $\left( \tilde{Z} - Z \right)$ with a coefficient $c$ to the
integrand of (\ref{Gamma 8 SG}) with the above replacements, we find that
the action density in terms of the rescaled parameters (\ref{rescaled
parameters}), with the 5-brane term omitted, becomes a constant
positive multiple of:
\begin{dmath}[compact, spread=3pt]
\label{action with field redefinition term} A^4 \tilde{B}^7 \left(
- \frac{42}{\tilde{B}^2} - \frac{7 \tilde{h}^2}{8 \tilde{B}^8} + \left(
\frac{1}{\tilde{B}^2} + \frac{\tilde{h}^2}{8 \tilde{B}^8} \right)^4 + c
\left( - \frac{19}{21 \tilde{B}^8} + \frac{89 \tilde{h}^2}{189
\tilde{B}^{14}} + \frac{23 \tilde{h}^4}{10368 \tilde{B}^{20}} + \frac{20395
\tilde{h}^6}{870912 \tilde{B}^{26}} + \frac{2225255 \tilde{h}^8}{3009871872
\tilde{B}^{32}} \right) \right)
\end{dmath}
Deriving the
field equations, one finds that a flat $\mathbf{R}^4$ times $\bar{H}^7$
solution exists for $c < c_{\mathrm{\max}} \equiv - \frac{734832}{2225255}
\simeq - 0.3302$. \ Thus a solution exists for precisely the range of $c$
for which the coefficient of the highest power of $\frac{1}{\tilde{B}}$ in
the action is negative. \ To study the field
equations, it is convenient to define
$x \equiv \frac{\tilde{h}}{\tilde{B}^3} = \frac{h}{B^3}$. \ The equations
then become two
simultaneous linear equations for $c$ and $\tilde{B}^6$. \ Solving the
equations one finds a hyperbola-like relation between $c$ and $x$, with $x
\rightarrow + \infty$ as $c \rightarrow c_{\mathrm{\max}}$ from below, and $c
\rightarrow - \infty$ as $x \rightarrow x_{\mathrm{\min}} \simeq 0.9364$ from
above. \ The range of $x$ perceived as the ``corner'' of the hyperbola
depends on the range of $x$ for which the curve is plotted, but when the
lower limit of the plot is about $x = 3$, so that $c$ does not drop below
about $2 c_{\mathrm{\max}}$, the ``corner'' of the hyperbola is from about
$x = 3$ to about $x = 8$.
One possibility is that this solution is unphysical, and either has tachyonic
modes or is subject to very large corrections at higher orders in $\kappa^{2 /
3}_{11}$. \ Another possibility is that there might be a ``field redefinition
group'' similar to the renormalization group, and that at low orders of
perturbation theory, we should use ``coordinates in field space'' best suited
to the geometry being studied. \ Doing that would minimize the size of the
higher-order corrections. \ Sensitivity to field redefinitions should decrease
as higher-order corrections are included, so we should use the \emph{principle
of minimal sensitivity} {\cite{Stephenson}} to choose the best field
redefinition at low orders of perturbation theory, as in perturbative
QCD. \ $c$ is the coefficient of the difference from the identity of a
perturbative field redefinition, that we are neglecting terms of quadratic and
higher order in the Taylor expansion of the action with respect to, so the
magnitude of $c$ should also be as small as possible. Thus we should
choose $c$ at the ``corner'' of the hyperbola in the $\left( x, c \right)$
plane. \ Choosing $x = 5.5$, we find $c \simeq - 0.38$, and $\tilde{B} \simeq
0.91$, which corresponds to
\begin{equation}
\label{best value of B} B \simeq 0.43 \kappa^{2 / 9}_{11}
\end{equation}
as the best
value of the radius of curvature of the $\bar{H}^7$. \ The best value of $h =
B^3 x$ is then $h \simeq 0.45 \kappa^{2 / 3}_{11}$.
The first Pontryagin class of an $\bar{H}^7$ is 0, so
the fluxes $H_{i j k l}$ wrapping individual 4-cycles of the
$\bar{H}^7$\hspace{-2pt} are constrained\hspace{-1pt} by\hspace{-1pt} flux
quantization such that
the integral of $\frac{1}{4!} H_{i j k l} dx^i \wedge dx^j \wedge dx^k \wedge
dx^l$ over a 4-cycle is equal to $\frac{2 \pi}{T_2}$ times an integer, where
$T_2 = \frac{1}{2} \left( \frac{4 \pi}{\kappa_{11}} \right)^{2 / 3}$ is the
fundamental membrane tension
{\cite{Rohm Witten, Witten flux quantization, Bousso Polchinski}}. \ If the
average flux $h$ is not exactly equal to the value required by the
solution of the field equations with flat extended dimensions, then the
extended dimensions will have to be slightly curved to compensate. \ With
$\kappa^{2 / 9}_{11}$ around an inverse TeV, thus around $10^{- 19}$ metres,
the observed de Sitter radius of the extended dimensions,
which is $16.0$ Gyr $= \hspace{1pt}
\hspace{1pt} 1.51 \times 10^{26}$ metres $= \hspace{1pt} \hspace{1pt}
\hspace{2pt} 0.94 \times 10^{61} \sqrt{G_N}$ {\cite{de Sitter radius}},
is around $10^{45} \kappa^{2 /
9}_{11}$, so the curvature of the extended dimensions is around $10^{- 90}
\kappa^{- 4 / 9}_{11}$. \ For an $\bar{H}^7$ with intrinsic volume
$\bar{V}_7$ around $10^{34}$, typical 4-cycles of the $\bar{H}^7$ also have
intrinsic 4-area
around $10^{34}$, so individual flux intensities can be adjusted to a relative
precision of around $10^{- 34}$. \ By choosing incommensurate flux numbers,
the average intensity of the fluxes wrapping any three 4-cycles can be
adjusted to a relative precision of around $10^{- 102}$. \ The number of
independent fluxes is the number of linearly independent harmonic 4-forms,
which is the 4th Betti number, which from the discussion on the pages before
and after (\ref{generalized Gauss Bonnet theorem}), near the end of
section~\ref{Introduction}, is likely to be roughly in proportion to the
intrinsic volume, up to a logarithmic correction factor {\cite{Gromov Volume
Bounded Cohomology, Xue, Clair Whyte}}. \ Thus flux
quantization does not prevent the observed de Sitter radius from being
attained.
The second derivative of (\ref{action with field redefinition term}) with
respect to $\tilde{B}$, at the above solution, is $\simeq - 19000
\tilde{A}^4$, so a deviation of $\tilde{B}$ from its value $\tilde{B} \simeq
0.91$ at the solution has positive energy, and this would normally correspond
to positive mass$^2$ for the mode where the deviation of $\tilde{B}$ from
$0.91$ depends on position in the 4 extended dimensions. \ This mode is a
scalar on the extended dimensions, and its only position dependence is its
position dependence along the extended dimensions. \ For a first approximation
I shall take the derivative part of its kinetic term as given by the Einstein
action plus gauge-fixing terms. \ The Einstein action density, up to a
positive overall factor and a total
derivative term which I shall temporarily drop, is:
\begin{dmath}[compact, spread=3pt]
\label{spin 2 action density} \sqrt{- G} R = - \frac{1}{4} \sqrt{- G} G^{I
J} G^{K L} G^{M N} \left( \partial_I G_{K M} \partial_J G_{L N} - 2
\partial_I G_{N K} \partial_M G_{J L} + 2 \partial_I G_{K L} \partial_M G_{N
J} - \partial_I G_{K L} \partial_J G_{M N} \right)
\end{dmath}
Extracting the part of this which has $\partial_{\mu}$ derivatives on $B$ in
the metric ansatz (\ref{metric ansatz for H7}), on page \pageref{metric ansatz
for H7}, we see that only the first and last terms in the right-hand side of
(\ref{spin 2 action density}) contribute, giving
\begin{equation}
\label{derivative part of kinetic term}- \frac{1}{4} \sqrt{- G}
\frac{1}{B^2} g^{kl} g^{mn} \left( g_{k m} g_{l n} - g_{kl} g_{mn} \right)
\partial_{\mu} B \partial^{\mu} B = \frac{21}{2} \sqrt{- G} \frac{1}{B^2}
\partial_{\mu} B \partial^{\mu} B.
\end{equation}
Thus this is a mode for which the
kinetic term coming from the Einstein action has the wrong sign, and its
positive mass$^2$ could potentially lead to exponential instead of oscillatory
time dependence. \ However a gauge-fixing term such as
\begin{equation}
\label{gauge fixing term}\frac{1}{2 \alpha}
\sqrt{- G} G^{I J} G^{K L} \left( \partial_K \varphi_{L I} - \beta \partial_I
\varphi_{K L} \right) G^{M N} \left( \partial_M \varphi_{N J} - \beta
\partial_J \varphi_{M N} \right),
\end{equation}
where $\varphi_{I J}$ is the perturbation
of $G_{I J}$ from the solution found above, and any value of $\beta$ except
$\beta = 1$ is allowed, will generically alter the derivative part of the
kinetic term of this mode, and can change its sign. \ ($\beta = 1$ is not
allowed because $G^{K L} \left( \partial_K \varphi_{L I} - \partial_I
\varphi_{K L} \right)$ has a residual invariance under $\varphi_{I J}
\rightarrow \varphi_{I J} + \partial_I \partial_J \lambda$, and can thus not
be set equal to an arbitrary 11-vector by a gauge transformation $\varphi_{I
J} \rightarrow \varphi_{I J} + D_I \varepsilon_J + D_J \varepsilon_I$.) \ Thus
no exponential time dependence of this mode is observable.
\section{The HW boundary}
\label{The HW boundary}
Assuming provisionally that the solution found in the previous subsection is
valid, I shall now show that it might be possible to introduce an HW boundary,
with the region around the boundary treated in a consistent semiclassical
approximation. \ This can at best be only semi-quantitative, because no higher
derivative corrections to the boundary conditions, corresponding to the
leading quantum corrections used in the bulk, are introduced. \ However
higher derivative boundary conditions have multiple solutions, most of which
will be unstable, with tachyonic excitations. \ We have to select the solution
free of tachyons, which will be the smoothest one, for which the higher
derivative corrections have the least possible significance. \ Thus it is
useful to find a solution that satisfies the semiclassical boundary conditions
(\ref{Israel boundary conditions}) and (\ref{bc for H}), since it should
provide a good starting point for finding the correct solution to the boundary
conditions with the higher derivative corrections included. \ I shall assume
that the
boundary condition (\ref{bc for H}) for the fluxes can be satisfied in an
average sense, and consider the boundary condition (\ref{Israel boundary
conditions}) for the metric.
For simplicity I shall consider a boundary whose topology is a smooth compact
quotient $\bar{H}^6$ of $H^6$, although it is not known whether $\bar{H}^6$ of
suffiently small intrinsic volume $\bar{V}_6$ actually exist. \ Better choices
for the boundary topology might be the Cartesian product of two small
intrinsic volume $\bar{H}^3$'s, or an Anderson 6-manifold {\cite{Anderson}},
but it does not appear to be known whether an $\bar{H}^7$ can have
minimal-area 6-cycles of these topologies. \ I shall assume that the boundary
is somewhere near a minimal-area 6-cycle of the closed $\bar{H}^7$ that is cut
into two disconnected parts along the boundary, and use the notation of
section \ref{Horava Witten theory}, starting on page \pageref{Horava Witten
theory}. \ The metric in this region has the form:
\begin{equation}
\label{metric near boundary} ds_{11}^2 = G_{IJ} dx^I dx^J = a \left( y
\right)^2 \eta_{\mu \nu} dx^{\mu} dx^{\nu} + b \left( y \right)^2 h_{a b}
dx^a dx^a + dy^2
\end{equation}
Here $a \left( y \right)$ is as in (\ref{metric ansatz}), on page
\pageref{metric ansatz}, so $a \left( y
\right) \rightarrow A$ away from the boundary, and $a \left( y \right)
\rightarrow 1$ on the boundary. \ Indices $a, b, c, \ldots$ are tangential to
the boundary, so that $x^A$ in (\ref{metric ansatz}) is now $\left( x^a, y
\right)$, and $h_{a b}$ is a metric of sectional curvature $- 1$.
\vspace{0.2cm}
The nonvanishing Riemann tensor components for the metric ansatz (\ref{metric
near boundary}) are:
\[ R_{\mu \nu \sigma \tau} = \frac{\dot{a}^2}{a^2} \left( G_{\mu \tau} G_{\nu
\sigma} - G_{\mu \sigma} G_{\nu \tau} \right), \]
\[ R_{a b c d} = b^2 R_{a b c d} \left( h \right) + \frac{\dot{b}^2}{b^2}
\left( G_{a d} G_{b c} - G_{a c} G_{b d} \right), \]
\begin{equation}
\label{Riemann tensor components} R_{\mu a \nu b} = - \frac{\dot{a}
\dot{b}}{ab} G_{\mu \nu} G_{a b}, \hspace{1.2cm} R_{\mu y \nu y} = -
\frac{\ddot{a}}{a} G_{\mu \nu}, \hspace{1.2cm} R_{ayby} = -
\frac{\ddot{b}}{b} G_{a b},
\end{equation}
together with the components related to these by the antisymmetries of the
Riemann tensor under the interchange of its first two indices, and under the
interchange of its last two indices, where a dot denotes
differentiation with respect to $y$, and $R_{a b c d} \left( h \right) =
h_{a d} h_{b c} - h_{a c} h_{b d}$
denotes the Riemann tensor calculated from the six-dimensional metric $h_{a
b}$.
\vspace{0.2cm}
Thus the nonvanishing Ricci tensor components for the metric ansatz
(\ref{metric near boundary}) are:
\[ R_{\nu \tau} = - 3 \frac{\dot{a}^2}{a^2} G_{\nu \tau} - 6 \frac{\dot{a}
\dot{b}}{ab} G_{\nu \tau} - \frac{\ddot{a}}{a} G_{\nu \tau} = - \left( 3
\frac{\dot{a}^2}{a^2} + 6 \frac{\dot{a} \dot{b}}{ab} + \frac{\ddot{a}}{a}
\right) G_{\nu \tau} \]
\[ R_{b d} = R_{b d} \left( h \right) - 5 \frac{\dot{b}^2}{b^2} G_{b d} - 4
\frac{\dot{a} \dot{b}}{ab} G_{b d} - \frac{\ddot{b}}{b} G_{b d} = - \left(
\frac{5}{b^2} + 5 \frac{\dot{b}^2}{b^2} + 4 \frac{\dot{a} \dot{b}}{ab} +
\frac{\ddot{b}}{b} \right) G_{b d} \]
\begin{equation}
\label{Ricci tensor components} R_{y y} = - 4 \frac{\ddot{a}}{a} - 6
\frac{\ddot{b}}{b}
\end{equation}
where I used that $R_{b d} \left( h \right) = - 5 h_{b d}$.
\vspace{1.0cm}
I shall look for a solution where the metric in the region of the boundary is
a small perturbation of what it would have been if the boundary was not there.
\ In the absence of the boundary $a$ is equal to the constant $A$, and $R_{A B
C D}$ is $\frac{1}{B^2} \left( G_{A D} G_{B C} - G_{A C} G_{B D} \right)$, the
Riemann tensor on the closed $\bar{H}^7$, where
$B$ has the value $B \simeq 0.43
\kappa^{2 / 9}_{11}$ found in the previous section. \ Thus from the second
line of (\ref{Riemann tensor components}), we require $\left( \frac{1}{b^2} +
\frac{\dot{b}^2}{b^2} \right) = \frac{1}{B^2}$, and assuming $b \left( y
\right)$ has its minimum value at $y = 0$, the solution of this is $b = B
\mathrm{\cosh} \left( \frac{y}{B} \right)$. \ This then also gives the correct
value $R_{ayby} = - \frac{1}{B^2} G_{a b}$, from the third line of
(\ref{Riemann tensor components}). \ I shall now work out what the effective
energy-momentum tensor components must be, in order for this form of $b \left(
y \right)$, with constant $a\left( y\right) = A$, to solve the classical
Einstein equations.
I assume the effective energy-momentum tensor $T_{I J}$ in the absence of the
boundary has the form:
\begin{equation}
\label{effective energy momentum tensor} T_{\mu \nu} = t^{\left( 1 \right)}
\left( y \right) G_{\mu \nu}, \hspace{1cm} T_{a b} = t^{\left( 2 \right)}
\left( y \right) G_{a b}, \hspace{1cm} T_{y y} = t^{\left( 3 \right)} \left(
y \right)
\end{equation}
The Einstein equations can be written:
\begin{equation}
\label{effective Einstein equations} R_{I J} + \kappa^2_{11} \left( T_{I J}
- \frac{1}{9} G_{I J} G^{K L} T_{K L} \right) = 0,
\end{equation}
and using the Ricci tensor components from (\ref{Ricci tensor components}),
these are:
\begin{equation}
\label{first Einstein eqn} \frac{\ddot{a}}{a} + 3 \frac{\dot{a}^2}{a^2} + 6
\frac{\dot{a} \dot{b}}{ab} + \frac{\kappa^2_{11}}{9} \left( 5 t^{\left( 1
\right)} \left( y \right) - 6 t^{\left( 2 \right)} \left( y \right) -
t^{\left( 3 \right)} \left( y \right) \right) = 0
\end{equation}
\begin{equation}
\label{second Einstein eqn} \frac{\ddot{b}}{b} + 5 \frac{\dot{b}^2}{b^2} + 4
\frac{\dot{a} \dot{b}}{ab} + \frac{5 \lambda}{b^2}
+ \frac{\kappa^2_{11}}{9} \left(
- 4 t^{\left( 1 \right)} \left( y \right) + 3 t^{\left( 2 \right)} \left( y
\right) - t^{\left( 3 \right)} \left( y \right) \right) = 0
\end{equation}
\begin{equation}
\label{third Einstein eqn} 4 \frac{\ddot{a}}{a} + 6 \frac{\ddot{b}}{b} +
\frac{\kappa^2_{11}}{9} \left( - 4 t^{\left( 1 \right)} \left( y \right) - 6
t^{\left( 2 \right)} \left( y \right) + 8 t^{\left( 3 \right)} \left( y
\right) \right) = 0
\end{equation}
In (\ref{second Einstein eqn}) I have introduced a factor $\lambda$
multiplying the $\frac{5}{b^2}$ term as a rough way of allowing for the
possibility that the boundary might be, for example, the Cartesian product of
two $\bar{H}^3$'s, or an Anderson 6-manifold {\cite{Anderson}}. \ Setting
$\lambda = 1$ gives the equation that follows from the Ricci tensor
components (\ref{Ricci tensor components}).
Eliminating the double derivatives between the three Einstein equations, we
find:
\begin{equation}
\label{no double derivatives} \frac{\dot{a}^2}{a^2} + 4 \frac{\dot{a}
\dot{b}}{ab} + \frac{5 \dot{b}^2}{2 b^2}
+ \frac{5 \lambda}{2 b^2} - \frac{1}{6}
\kappa^2_{11} t^{\left( 3 \right)} = 0,
\end{equation}
hence:
\begin{equation}
\label{eqn for a dot} \frac{\dot{a}}{a} = - 2 \frac{\dot{b}}{b} +
\frac{1}{2} \sqrt{6 \frac{\dot{b}^2}{b^2} - \frac{10 \lambda}{b^2} +
\frac{2}{3} \kappa_{11}^2 t^{\left( 3 \right)}},
\end{equation}
where I assumed that $\dot{b}$ does not change sign between the boundary and
the main part of the bulk, so the sign of the square root follows from
$\dot{a} = 0$ and the convention stated at the start of section \ref{Horava
Witten theory}, on page \pageref{Horava Witten theory}, that $y > y_1$ in the
bulk, where $y_1$ is the value of $y$ at the boundary.
The second Einstein equation, (\ref{second Einstein eqn}), now becomes:
\begin{equation}
\label{eqn for b double dot} \frac{\ddot{b}}{b} - 3 \frac{\dot{b}^2}{b^2} +
2 \frac{\dot{b}}{b} \sqrt{6 \frac{\dot{b}^2}{b^2} - \frac{10 \lambda}{b^2} +
\frac{2}{3} \kappa_{11}^2 t^{\left( 3 \right)}} + \frac{5 \lambda}{b^2} +
\frac{\kappa_{11}^2}{9} \left( - 4 t^{\left( 1 \right)}
+ 3 t^{\left( 2 \right)}
- t^{\left( 3 \right)} \right) = 0
\end{equation}
From subsection 2.4 of {\cite{CCHT}}, a solution of (\ref{eqn for a dot}) and
(\ref{eqn for b double dot}) solves all three Einstein equations, provided the
energy-momentum tensor is conserved, and the square root does not vanish
identically for the solution. \ The square root does not vanish for the
solution I will consider, so it is sufficient to consider (\ref{eqn for a
dot}) and (\ref{eqn for b double dot}).
Requiring that (\ref{eqn for a dot}) and (\ref{eqn for b double dot}) are
solved by $a \left( y \right) = A$, $b \left( y \right) = B \mathrm{\cosh}
\frac{y}{B}$, we find the effective energy-momentum tensor components:
\begin{equation}
\label{t upper 3} \frac{2}{3} \kappa_{11}^2 t^{\left( 3 \right)} =
\frac{10\, \mathrm{\sinh}^2 \frac{y}{B}}{B^2
\mathrm{\cosh}^2 \frac{y}{B}} + \frac{10 \lambda}{B^2 \mathrm{\cosh}^2
\frac{y}{B}}
\end{equation}
\begin{equation}
\label{t1 t2 t3} \frac{\kappa_{11}^2}{9} \left( - 4 t^{\left( 1 \right)} + 3
t^{\left( 2 \right)} - t^{\left( 3 \right)} \right) = - \left( \frac{1}{B^2}
+ 5 \frac{\mathrm{\sinh}^2 \frac{y}{B}}{B^2 \mathrm{\cosh}^2 \frac{y}{B}} +
\frac{5 \lambda}{B^2 \mathrm{\cosh}^2 \frac{y}{B}} \right)
\end{equation}
Substituting in these values of the $t^{\left( i \right)}$, the Einstein
equations (\ref{eqn for a dot}) and (\ref{eqn for b double dot}) become:
\begin{equation}
\label{first eqn in background} \frac{\dot{a}}{a} = - 2 \frac{\dot{b}}{b} +
\frac{1}{2} \sqrt{6 \frac{\dot{b}^2}{b^2} - \frac{10 \lambda}{b^2} +
\frac{10 \mathrm{\sinh}^2 \frac{y}{B}}{B^2 \mathrm{\cosh}^2 \frac{y}{B}} +
\frac{10 \lambda}{B^2 \mathrm{\cosh}^2 \frac{y}{B}}}
\end{equation}
\begin{dmath}[compact, spread=3pt]
\label{2nd eqn in background} \frac{\ddot{b}}{b} - 3 \frac{\dot{b}^2}{b^2} +
2 \frac{\dot{b}}{b} \sqrt{6 \frac{\dot{b}^2}{b^2} - \frac{10 \lambda}{b^2} +
\frac{10 \mathrm{\sinh}^2 \frac{y}{B}}{B^2 \mathrm{\cosh}^2 \frac{y}{B}} +
\frac{10 \lambda}{B^2 \mathrm{\cosh}^2 \frac{y}{B}}} + \frac{5
\lambda}{b^2} - \left( \frac{1}{B^2} + 5 \frac{\mathrm{\sinh}^2
\frac{y}{B}}{B^2 \mathrm{\cosh}^2 \frac{y}{B}} + \frac{5 \lambda}{B^2
\mathrm{\cosh}^2 \frac{y}{B}} \right) = 0
\end{dmath}
We now let $a\left(y\right) = \left( 1 + p\left(y\right) \right) A$,
$b\left(y\right) = \left( 1 + q\left(y\right) \right) B
\mathrm{\cosh} \frac{y}{B}$, and expand to first order in $p$ and $q$. \ Then
(\ref{first eqn in background}) and (\ref{2nd eqn in background}) become:
\begin{equation}
\label{p dot eqn} \dot{p} = - \frac{5}{4} \dot{q} + \frac{5 \lambda q}{4 B
\mathrm{\sinh} \frac{y}{B} \mathrm{\cosh} \frac{y}{B}}
\end{equation}
\begin{equation}
\label{q double dot eqn} \ddot{q} + 7 \dot{q} \frac{\mathrm{\sinh}
\frac{y}{B}}{B \mathrm{\cosh} \frac{y}{B}} - \frac{5 \lambda q}{B^2
\mathrm{\cosh}^2 \frac{y}{B}} = 0
\end{equation}
For a first estimate of the boundary conditions (\ref{Israel boundary
conditions}), on page \pageref{Israel boundary conditions}, I shall neglect
the flux terms in $\bar{R}_{U V W X}$, defined
in (\ref{R bar}), by assuming, if necessary, that $H_{\hat{y} U V W}$ is
smaller than its average value, near the boundary. \ The energy-momentum
tensor on the boundary is then:
\begin{dmath}[compact, spread=3pt]
\label{energy momentum tensor on boundary} \tilde{T}_{U V} = \frac{1}{16 \pi
\kappa^2_{11}} \left( \frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left(
\frac{4}{30} \mathrm{{tr}} F_{U W} F_V \, \!^W - \frac{1}{30}
\tilde{G}_{U V} \mathrm{{tr}} F_{W X} F^{W X} - 2 \bar{R}_{U W X Y}
\bar{R}_V \, \!^{W X Y} + \frac{1}{2} \tilde{G}_{U V} \bar{R}_{W X Y Z}
\bar{R}^{W X Y Z} \right)
\end{dmath}
I shall work in the leading order of the Lukas-Ovrut-Waldram {\cite{Lukas
Ovrut Waldram}} harmonic expansion of the energy-momentum tensor (\ref{energy
momentum tensor on boundary}) on the boundary, and assume that
$\mathrm{{tr}} F_{a c} F_b \, \!^c$ is a multiple of $\tilde{G}_{a b}$,
and also that $\bar{R}_{a c d e} \bar{R}_b \, \!^{c d e}$ is a multiple of
$\tilde{G}_{a b}$, which it is if the boundary is an $\bar{H}^6$, but is not
if the boundary is an $\bar{H}^3 \times \bar{H}^3$, for example. \ We then
find:
\begin{equation}
\label{T tilde mu nu} \tilde{T}_{\mu \nu} = \frac{1}{16 \pi \kappa^2_{11}}
\left( \frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left( - \frac{1}{30}
\mathrm{{tr}} F_{c d} F^{c d} + \frac{1}{2} \bar{R}_{c d e f}
\bar{R}^{c d e f} \right) \tilde{G}_{\mu \nu}
\end{equation}
\begin{equation}
\label{T tilde a b} \tilde{T}_{a b} = \frac{1}{16 \pi \kappa^2_{11}} \left(
\frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left( - \frac{1}{90}
\mathrm{{tr}} F_{c d} F^{c d} + \frac{1}{6} \bar{R}_{c d e f}
\bar{R}^{c d e f} \right) \tilde{G}_{a b}
\end{equation}
We define $\tilde{t}^{\left( 1 \right)}$ and $\tilde{t}^{\left( 2 \right)}$
by:
\begin{equation}
\label{definitions of T tildes} \tilde{T}_{\mu \nu} = \tilde{t}^{\left( 1
\right)} \tilde{G}_{\mu \nu}, \hspace{1.8cm} \tilde{T}_{ab}
= \tilde{t}^{\left( 2 \right)} \tilde{G}_{ab}
\end{equation}
Then by subsection 2.3.9 of {\cite{CCHT}}, the boundary conditions for
$a\left(y\right)$ and $b\left(y\right)$ are:
\begin{equation}
\left. \left. \label{bcs in terms of t tildes} \frac{\dot{a}}{a} \right|_{y
= y_{1 +}} = \frac{\kappa_{11}^2}{18} \left( - 5 \tilde{t}^{\left( 1
\right)} + 6 \tilde{t}^{\left( 2 \right)} \right), \hspace{6.0ex}
\frac{\dot{b}}{b} \right|_{y = y_{1 +}} = \frac{\kappa_{11}^2}{18} \left( 4
\tilde{t}^{\left( 1 \right)} - 3 \tilde{t}^{\left( 2 \right)} \right)
\end{equation}
Thus:
\begin{equation}
\left. \label{final bc for a} \frac{\dot{a}}{a} \right|_{y = y_{1 +}} = -
\frac{1}{96 \pi} \left( \frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left( -
\frac{1}{30} \mathrm{{tr}} F_{c d} F^{c d} + \frac{1}{2} \bar{R}_{c d e
f} \bar{R}^{c d e f} \right)
\end{equation}
\begin{equation}
\left. \label{final bc for b} \frac{\dot{b}}{b} \right|_{y = y_{1 +}} =
\frac{1}{96 \pi} \left( \frac{\kappa_{11}}{4 \pi} \right)^{2 / 3} \left( -
\frac{1}{30} \mathrm{{tr}} F_{c d} F^{c d} + \frac{1}{2} \bar{R}_{c d e
f} \bar{R}^{c d e f} \right)
\end{equation}
The vacuum Yang-Mills fields on the boundary are quantized, for example by a
Dirac
quantization condition as in subsection 5.3 of {\cite{CCHT}} if they are in
the Cartan subalgebra of $E_8$, so the right-hand sides of (\ref{final bc for
a}) and (\ref{final bc for b}) are proportional to $b^{- 4}_1$, where $b_1
\equiv b \left( y_1 \right)$ is the value of $b$ on the boundary. \ Moss's
derivation of the $R_{U V W X} R^{U V W X}$ term in (\ref{Yang Mills term})
used an expansion scheme in which Ricci tensor and scalar terms, if present,
would only show up at higher orders. \ If the $R_{U V W X} R^{U V W X}$ term
was in fact the first term in a Lovelock-Gauss-Bonnet term of the form $R_{U V
W X} R^{U V W X} - 4 R_{U V} R^{U V} + R^2$, the size of the curvature terms
in (\ref{final bc for a}) and (\ref{final bc for b}) would be increased by a
factor of 6 if the boundary is an $\bar{H}^6$, with the main contribution
coming from the square of the Ricci scalar. \ This would make it easier to
keep the right-hand side of (\ref{final bc for a}) negative and the right-hand
side of (\ref{final bc for b}) positive, while introducing enough vacuum
Yang-Mills fields to break $E_8$ to $\mathrm{{SU}} \left( 3 \right)
\times \mathrm{{SU}} \left( 2 \right) \times \mathrm{U} \left( 1
\right)_Y$ and produce the chiral fermions.
To find the solution of (\ref{p dot eqn}) and (\ref{q double dot eqn}) such
that $p$ and $q$ tend to 0 as $y \rightarrow \infty$, we define
$\mathrm{\tanh} \frac{y}{B} = x$, so that $x \rightarrow 1$ as
$y \rightarrow \infty$. \
The equations then become:
\begin{equation}
\label{p eqn with x} \frac{dp}{dx} = - \frac{5}{4} \frac{dq}{dx} + \frac{5
\lambda q}{4 x}
\end{equation}
\begin{equation}
\label{q eqn with x} \left( 1 - x^2 \right) \frac{d^2 q}{dx^2} + 5 x
\frac{dq}{dx} - 5 \lambda q = 0
\end{equation}
For (\ref{q eqn with x}), we try $q \rightarrow \left( 1 - x \right)^{\alpha}$
near $x = 1$, which gives $\alpha = \frac{7}{2}$. \ We then find the
expansion:
\begin{dmath}[compact, spread=3pt]
\label{expansion of q} q = \left( 1 - x \right)^{\frac{7}{2}} + \frac{\left(
20 \lambda - 35 \right) \left( 1 - x \right)^{\frac{9}{2}}}{36} +
\frac{\left( 400 \lambda^2 - 1240 \lambda + 945 \right) \left( 1 - x
\right)^{\frac{11}{2}}}{3168} + \frac{\left( 8000 \lambda^3 - 29200
\lambda^2 + 32540 \lambda - 10395 \right) \left( 1 - x
\right)^{\frac{13}{2}}}{494208} + \cdots
\end{dmath}
For $\lambda = 1$, (\ref{expansion of q}) looks qualitatively like the base of
a parabola centred at $x = 1$, and is $\simeq 0.6184$ for $x = 0$.
The solution of (\ref{p eqn with x}) is $p = - \frac{5}{4} q - \frac{5
\lambda}{4} \int^1_x \frac{q \left( x' \right)}{x'} dx'$. \ Thus $p$ is $\leq
0$ for $0 < x \leq 1$, and has a logarithmic singularity as $x \rightarrow 0$.
We now try to satisfy the boundary conditions (\ref{final bc for a}) and
(\ref{final bc for b}) with $a = \left( 1 + kp \right) A$, $b = \left( 1 + kq
\right) B \mathrm{\cosh} \frac{y}{B}$, where $p$ and $q$ are the solutions
just found, and $k$ is a constant.
We write the boundary conditions (\ref{final bc for a}) and (\ref{final bc for
b}) as:
\begin{equation}
\label{boundary conditions in terms of rho} \left. \frac{\dot{a}}{a}
\right|_{y = y_{1 +}} = - \rho \frac{\kappa^{2 / 3}_{11}}{b^4_1},
\hspace{3.2cm} \left. \frac{\dot{b}}{b} \right|_{y = y_{1 +}} = \rho
\frac{\kappa^{2 / 3}_{11}}{b^4_1},
\end{equation}
where the number $\rho$ is:
\begin{equation}
\label{definition of rho} \rho \equiv \frac{1}{\bar{V}_6} \int_{\beta} d^6
\tilde{x} \sqrt{h} \frac{1}{96 \pi \left( 4 \pi \right)^{2 / 3}} h^{a c}
h^{b d} \left( - \frac{1}{30} \mathrm{{tr}} F_{a b} F_{c d} +
\frac{1}{2} \bar{R}_{a b} \, \!^e \, \!_f \bar{R}_{c d e} \, \!^f \right),
\end{equation}
where $\bar{V}_6 \equiv \int_{\beta} d^6 \tilde{x} \sqrt{h}$ is the intrinsic
volume of the boundary $\beta$ in the metric $h_{a b}$, which if the boundary
is an $\bar{H}^6$ is defined to have sectional curvature $- 1$. \ If the
boundary is an $\bar{H}^6$, with $\bar{R}_{a b} \, \!^e \, \!_f = h_{a f}
\delta_b \, \!^e - h_{b f} \delta_a \, \!^e$, then the Riemann$^2$ term in
$\rho$ is $\frac{60}{192 \pi \left( 4 \pi \right)^{2 / 3}} \simeq 0.01840$. \
If the spin connection is naturally embedded in the $E_8$ Yang-Mills gauge
group {\cite{Wilczek embedding spin connection, Charap Duff, Witten Shelter
Island, Candelas Horowitz Strominger Witten}}, which by the Bianchi identity
for the Riemann tensor gives a solution of the classical Yang-Mills equations
if $h_{a b}$ is an Einstein metric, then the Yang-Mills term in $\rho$ is $-
2$ times the Riemann$^2$ term, so $\rho$ is negative.
I shall assume that for moderate values of $\bar{V}_6$ it is possible to
introduce vacuum gauge fields in the $E_8$ Cartan subalgebra, topologically
stabilized by a Dirac quantization condition as in subsection 5.3 of
{\cite{CCHT}}, with field strengths proportional to 4, 5, or 6 of a larger
number of linearly independent harmonic 2-forms on the boundary, such that
$E_8$ is broken to $\mathrm{{SU}} \left( 3 \right) \times
\mathrm{{SU}} \left( 2 \right) \times \mathrm{U} \left( 1 \right)_Y$ and
the Standard Model chiral fermion zero modes are produced, and the Yang-Mills
term in $\rho$ is sufficiently ``diluted'' that $\rho$ is \emph{positive}.
\ If the Riemann$^2$ term in (\ref{Yang Mills term}) is the first term of the
Lovelock-Gauss-Bonnet combination, then the Riemann and Ricci contribution to
$\rho$ for $\bar{H}^6$ would be 6 times larger, so $\rho$ would be positive
even if the spin connection was embedded in the gauge group.
With $\mathrm{\tanh} \frac{y}{B} = x$ as before, and $x_1$ denoting the value
of $x$ at the boundary, the sum of the boundary conditions (\ref{boundary
conditions in terms of rho}) gives:
\begin{equation}
\label{equation determining k} k \left( 1 - x_1^2 \right) \left( -
\frac{1}{4} \left. \frac{dq}{dx} \right|_{x = x_1} + \frac{5 q \left( x_1
\right)}{4 x_1} \right) + x_1 = 0,
\end{equation}
where (\ref{p eqn with x}) has been used with $\lambda = 1$. \ The logarithmic
singularity of $p$ as $x \rightarrow 0_+$ means that we require $x_1 > 0$ for
the assumption that $\left| kp \right| \ll 1$ to be valid, so since
$\frac{dq}{dx} \leq 0$ and $q \geq 0$ for $0 \leq x \leq 1$, (\ref{equation
determining k}) implies that $k < 0$. \ Using $B \simeq 0.43 \kappa^{2 /
9}_{11}$, from (\ref{best value of B}), on page \pageref{best value of B}, the
second equation of (\ref{boundary conditions in terms of rho}) becomes:
\begin{equation}
\label{second bc in terms of x} k \left( 1 - x_1^2 \right) \left.
\frac{dq}{dx} \right|_{x = x_1} + x_1 \simeq \frac{201 \rho \left( 1 - 4 kq
\left( x_1 \right) \right)}{\left( \sqrt{\frac{1 + x_1}{1 - x_1}} +
\sqrt{\frac{1 - x_1}{1 + x_1}} \right)^4}
\end{equation}
Using (\ref{equation determining k}) to express $k$ in terms of $x_1$, we find
that $kq_1 \equiv kq \left( x_1 \right)$, as a function of $x_1$, looks
qualitatively like an upside-down parabola with a peak value of 0 at $x_1 =
0$, and $\simeq - 0.25$ at $x_1 \simeq 0.63$. \ Substituting for $k$ from
(\ref{equation determining k}), we find from (\ref{second bc in terms of x})
that for $\rho = 0.01840$, $x_1 \simeq 0.165$, for which we have $kp_1 \equiv
kp \left( x_1 \right) \simeq 0.043$, hence $A \simeq 0.96$, and $kq_1 \simeq -
0.020$, hence $b_1 \simeq 0.99 B \simeq 0.43 \kappa^{2 / 9}_{11}$. \ Thus
working to first order in $kp$ and $kq$ has been justified. \ For $\rho \simeq
0.11$, corresponding to an $\bar{H}^6$ boundary with no vacuum Yang-Mills
fields, if the Riemann$^2$ term in (\ref{Yang Mills term}) is the start of the
Lovelock-Gauss-Bonnet combination, we find $x_1 \simeq 0.498$, for which $kp_1
\simeq 0.24$, hence $A \simeq 0.81$, and $kq_1 \simeq - 0.16$, hence $b_1
\simeq 0.97 B \simeq 0.42 \kappa^{2 / 9}_{11}$.
From (\ref{equation determining k}) we find that $\frac{b_1}{B}$, as a
function of $x_1$, decreases smoothly from a peak value of 1 at $x_1 = 0$, to
a minimum value $\simeq 0.967$ at $x_1 \simeq 0.56$, which is around the
largest value of $x_1$ for which it is reasonable to work to first order in
$kp$ and $kq$.
The Giudice-Rattazzi-Wells (GRW) estimate of the expansion parameter for
graviton loop corrections in $11$ dimensions, in the sense that perturbation
theory must fail when the expansion parameter is $> 1$ {\cite{Giudice Rattazzi
Wells}}, is:
\begin{equation}
\label{GRW estimate} \frac{\kappa_{11}^2}{945 \left( 2 \pi \right)^4} \left(
\frac{E}{2 \pi} \right)^9 \simeq \left( 0.03 \kappa_{11}^{2 / 9} E
\right)^9,
\end{equation}
where $E$ is the relevant energy of the process. \ For a wave of wavelength $2
\pi b_1$, $E$ is $\frac{1}{b_1}$, so the GRW estimate suggests that $b_1$ must
be larger than around $0.03 \kappa_{11}^{2 / 9}$. \ Thus for $b_1 \simeq 0.42
\kappa^{2 / 9}_{11}$, this requirement is satisfied by a substantial margin.
From just before (\ref{Breit Wigner cross section}), on page
\pageref{Breit Wigner cross section}, the first hypothesis
considered in section \ref{Introduction} leads to the
expectation that $0 < A \leq 0.72$ if $\frac{3 A}{B}=1.8 \textrm{ TeV}$, in
order to satisfy the experimental constraint \cite{Franceschini et al}
that $\kappa_{11}^{-2/9} \geq
0.36\textrm{ TeV}$, with $B \simeq 0.43 \kappa_{11}^{2/9}$ from section
\ref{The bulk}, while from just before the start of section \ref{Horava Witten
theory}, if $\frac{A}{B}=1.8 \textrm{ TeV}$, then the requirement
that $A\leq 1$, which follows from $k < 0$, leads to $\kappa_{11}^{-2/9}\geq
0.78\textrm{ TeV}$, hence
$M_{11}\geq 3.3\textrm{ TeV}$.
To study the Kaluza-Klein modes of the supergravity multiplet we have to
expand $\Gamma^{\left( \mathrm{{bos}} \right)}_{\mathrm{{HW}}}$, in
(\ref{Gamma HW}), including $\Gamma^{\left( 8, \mathrm{{bos}}
\right)}_{\mathrm{{SG}}}$, in (\ref{Gamma 8 SG}), to quadratic order in
small fluctuations about the solution found above. \ For a first estimate I
shall instead consider a massless scalar field $\Phi$ in the bulk, which is
intended to represent a small fluctuation of a component of any of the
supergravity fields, and retain only its classical action. \ Dropping
initially also $R \Phi$ and $H^2 \Phi$ terms, the equation for the small
fluctuation $\Phi$ is then:
\begin{equation}
\label{small fluctuation equation} - \frac{1}{\sqrt{- G}} \partial_I \left(
\sqrt{- G} G^{I J} \partial_J \Phi \right) = 0
\end{equation}
Trying an ansatz $\Phi \left( x^{\mu}, x^a, y \right) = \varphi \left( x^{\mu}
\right) \psi \left( x^a, y \right)$, we find from (\ref{small fluctuation
equation}) that:
\begin{dmath}
\label{separation of variables} - \frac{a^2}{b^2 \psi \left( x^c, y \right)
\sqrt{h}} \partial_a \left( \sqrt{h} h^{a b} \partial_b \psi \left( x^c, y
\right) \right) - \frac{a^2}{a^4 b^6 \psi \left( x^c, y \right)} \partial_y
\left( a^4 b^6 \partial_y \psi \left( x^c, y \right) \right) =
\frac{1}{\varphi \left( x^{\mu} \right) \sqrt{- g}} \partial_{\mu} \left(
\sqrt{- g} g^{\mu \nu} \partial_{\nu} \varphi \left( x^{\mu} \right) \right)
\end{dmath}
The left-hand side of (\ref{separation of variables}) is independent of
$x^{\mu}$ and the right-hand side is independent of $x^a$ and $y$, hence each
side must be a constant. \ The left-hand side is a positive operator on a
compact manifold so must be a non-negative constant $\frac{A^2}{B^2}m^2 \geq
0$. \ Away from the neighbourhood of the small boundary, the bulk is a
compact hyperbolic 7-manifold $\bar{H}^7$ with intrinsic volume $\bar{V}_7
\sim 10^{34}$ and curvature radius $B \simeq 0.43 \kappa_{11}^{2/9}$, and
$a = A$, so for most of the low-lying KK modes of the bulk, the KK
masses will to a first approximation not be affected by the presence of the
small boundary. \ So for most of the low-lying KK modes of the $d=11$
supergravity multiplet, $m^2$ will to a first approximation be an intrinsic
eigenvalue of the appropriate quantum-corrected differential operator, for
example the negative of the quantum-corrected Laplace-Beltrami operator, on
the corresponding closed $\bar{H}^7$ with curvature radius 1, with the
leading quantum corrections following from (\ref{Gamma 8 SG}).
For the negative of the Laplace-Beltrami operator $\Delta$, with quantum
corrections neglected, the left-hand side for the lightest massive modes is
estimated from (\ref{lower bound on lambda 1}), on page \pageref{lower bound
on lambda 1}, as $\frac{\left( n - 1 \right)^2}{2^2} \frac{A^2}{B^2} =
9 \frac{A^2}{B^2}$. \ So we find:
\begin{equation}
\label{mass at boundary of lightest KK modes} - \frac{1}{\sqrt{- g}}
\partial_{\mu} \left( \sqrt{- g} g^{\mu \nu} \partial_{\nu} \varphi \left(
x^{\mu} \right) \right) + 9 \frac{A^2}{B^2} \varphi \left( x^{\mu} \right)
= 0
\end{equation}
Thus for the first hypothesis in section \ref{Introduction}, the mass of the
lightest massive gravitational KK modes, as seen on the boundary, is $3
\frac{A}{B}$.
Most of the modes in the bulk couple only with gravitational strength to the
fields on the boundary, because the values on the boundary of their normalized
wavefunctions are suppressed by the reciprocal of the square root of the large
volume of the bulk. \ However if, as suggested in section \ref{Introduction},
there are around $10^{30}$ approximately degenerate KK modes of the bulk with
masses around 1.8 TeV as seen on the boundary, the modes in the bump could
equivalently be a relatively small number of linear combinations of these
modes, whose normalized wavefunctions are large near the boundary and small
away from the boundary, and these modes would have correspondingly large
couplings to the fields on the boundary.
To keep the number $\rho$ defined in (\ref{definition of rho})
positive, while introducing enough
vacuum Yang-Mills fields to break $E_8$ to $\mathrm{{SU}} \left( 3
\right) \times \mathrm{{SU}} \left( 2 \right) \times \mathrm{U} \left( 1
\right)_Y$ and produce the chiral fermions, it might be helpful if the
boundary had a large enough intrinsic volume for the $E_8$ breaking and chiral
fermions to arise from quantized Yang-Mills fluxes that are ``dilute'', in
the sense that only a small proportion of the possible quantized Yang-Mills
fluxes are present, and these have flux numbers $\pm 1$. However if the
boundary is an $\bar{H}^6$, its intrinsic volume $\bar{V}_6$ is approximately
fixed by the value $\alpha_U$ of the QCD fine structure constant $\alpha_s =
\frac{g^2_s}{4 \pi}$ at unification, together with the fact that from
the discussion after (\ref{second bc in terms of x}), $b_1 = b\left(y_1
\right) \simeq B \simeq 0.43 \kappa_{11}^{2/9}$.
I shall assume that the Yang-Mills $E_8$ at $y_1$ is
broken to the Standard Model $\mathrm{{SU}} \left( 3 \right) \times
\mathrm{{SU}} \left( 2 \right) \times \mathrm{U} \left( 1 \right)_Y$ by
topologically stabilized vacuum gauge fields in the Cartan subalgebra of the
$E_8$. \ These were shown in subsection 5.3 of {\cite{CCHT}} to be
restricted by Dirac quantization conditions to lie on a lattice of isolated
points in the Cartan subalgebra. \ I shall assume that the unbroken
$\mathrm{{SU}} \left( 3 \right) \times \mathrm{{SU}} \left( 2
\right) \times \mathrm{U} \left( 1 \right)_Y$ is naturally embedded within an
$\mathrm{{SU}} \left( 9 \right)$ subgroup of $E_8$, and from subsection
5.2 of {\cite{CCHT}}, the $E_8$ generators $T^{\mathcal{A}}$ in (\ref{Yang
Mills term}) are normalized so
that the structure constants of the $\mathrm{{SU}} \left( 3 \right)$
subgroup have the same normalization as in the definition of $g_s$ in
subsection 9.1 of {\cite{PDG}}, corresponding to $\mathrm{{SU}} \left( 3
\right)$ generators $t^{\alpha}$ normalized to $\mathrm{{tr}} \left(
t^{\alpha} t^{\beta} \right) = \frac{1}{2} \delta^{\alpha \beta}$, so
performing the trace over the $E_8$ generators using $\mathrm{{tr}}
T^{\mathcal{A}} T^{\mathcal{B}} = 30 \delta^{\mathcal{A}\mathcal{B}}$, we find
that
\begin{equation}
\label{alpha U} \alpha_U = \frac{\kappa^2_{11}}{b^6_1 \bar{V}_6} \left(
\frac{4 \pi}{\kappa_{11}} \right)^{2 / 3} = \frac{5.405}{\bar{V}_6} \left(
\frac{\kappa^{2 / 9}_{11}}{b_1} \right)^6 \simeq \frac{860}{\bar{V}_6},
\end{equation}
where $\bar{V}_6$ is the volume of the boundary in the metric $h_{a b}$,
which is the intrinsic volume of the boundary, if the boundary is an
$\bar{H}^6$.
In the Dienes-Dudas-Gherghetta version of accelerated unification $\alpha_U
\simeq \frac{1}{52}$ {\cite{DDG1, DDG2}}, and in the Arkani-Hamed-Cohen-Georgi
version, $\alpha_U \simeq \frac{1}{24}$ {\cite{ACG}}. \
If the boundary is an $\bar{H}^6$, then by (\ref{alpha U}) these imply
respectively $\bar{V}_6 \simeq 45000$ and $\bar{V}_6 \simeq 21000$, which by
the generalized Gauss-Bonnet theorem (\ref{generalized Gauss Bonnet theorem}),
on page \pageref{generalized Gauss Bonnet theorem}, correspond respectively to
Euler numbers $\chi \left( \bar{H}^6 \right) \simeq - 2700$ and $\chi \left(
\bar{H}^6 \right) \simeq - 1300$.
\begin{figure}[t]
\setlength{\unitlength}{1cm}
\begin{picture}(12.0,4.05)
\thicklines
\put(2.39,3.67){$u_1$}
\put(2.51,3.28){\circle*{0.265}}
\put(2.51,3.4){\line(1,0){1.40}}
\put(2.51,3.29){\line(1,0){1.40}}
\put(2.51,3.18){\line(1,0){1.40}}
\put(3.82,3.67){$u_2$}
\put(3.94,3.28){\circle*{0.265}}
\put(4.05,3.29){\line(1,0){1.40}}
\put(5.39,3.67){$u_3$}
\put(5.51,3.28){\circle*{0.265}}
\put(5.56,3.29){\line(1,0){1.48}}
\put(7.05,3.67){$u_4$}
\put(7.17,3.28){\circle*{0.265}}
\put(7.25,3.29){\line(1,0){1.43}}
\put(8.72,3.67){$u_8$}
\put(8.84,3.28){\circle*{0.265}}
\put(8.84,3.29){\line(1,0){1.48}}
\put(8.84,3.23){\line(0,-1){1.56}}
\put(10.28,3.67){$u_7$}
\put(10.4,3.28){\circle*{0.265}}
\put(10.25,3.33){\line(-1,-1){1.47}}
\put(10.35,3.23){\line(-1,-1){1.50}}
\put(10.45,3.16){\line(-1,-1){1.50}}
\put(8.20,1.65){$u_6$}
\put(8.86,1.75){\circle*{0.265}}
\put(10.28,0.55){$u_9$}
\put(10.4,1.05){\circle*{0.265}}
\put(7.08,0.05){$u_5$}
\put(7.2,0.54){\circle*{0.265}}
\put(8.98,1.64){\line(5,-2){1.56}}
\put(10.41,3.25){\line(0,-1){2.28}}
\put(7.15,3.17){\line(0,-1){2.50}}
\multiput(7.33,0.54)(0.18,0.03){17}{\circle*{0.06}}
\put(8.45,0.18){$\frac{1+\sqrt{5}}{2}$}
\end{picture}
\caption{Coxeter diagram for the $d = 6$ hyperbolic Coxeter polytope $T_3$}
\label{Coxeter diagram for T3}
\end{figure}
In 4 dimensions the Davis manifold {\cite{Davis manifold}} with $\chi = 26$,
which is an orientable spin manifold {\cite{Ratcliffe Tschantz Davis
manifold}}, can be constructed from a hyperbolic Coxeter simplex by a variant
of Selberg's Lemma {\cite{Selberg, Everitt Maclachlan}}. \ There are no
hyperbolic Coxeter simplexes in 5 or more dimensions {\cite{Lanner}}, and no
hyperbolic Coxeter polytopes with $n + 2$ $\left( n - 1 \right)$-dimensional
faces in 6 or more dimensions {\cite{Kaplinskaya, Esselmann}}. \ There are
exactly 3 hyperbolic Coxeter polytopes with 9 5-dimensional faces in 6
dimensions {\cite{Tumarkin}}, and the Coxeter diagram for one of these, which
I will call $T_3$, is shown in Figure \ref{Coxeter diagram for T3}. \ Twice
the Gram matrix for $T_3$ is:
\begin{equation}
\label{twice T3 Gram matrix} \mathcal{G} \equiv \left(
\begin{array}{ccccccccc}
2 & - k & 0 & 0 & 0 & 0 & 0 & 0 & 0\\
- k & 2 & - 1 & 0 & 0 & 0 & 0 & 0 & 0\\
0 & - 1 & 2 & - 1 & 0 & 0 & 0 & 0 & 0\\
0 & 0 & - 1 & 2 & - 1 & 0 & 0 & - 1 & 0\\
0 & 0 & 0 & - 1 & 2 & 0 & 0 & 0 & - 2 k\\
0 & 0 & 0 & 0 & 0 & 2 & - k & - 1 & - 1\\
0 & 0 & 0 & 0 & 0 & - k & 2 & - 1 & - 1\\
0 & 0 & 0 & - 1 & 0 & - 1 & - 1 & 2 & 0\\
0 & 0 & 0 & 0 & - 2 k & - 1 & - 1 & 0 & 2
\end{array} \right),
\end{equation}
where $k \equiv \frac{1 + \sqrt{5}}{2} = 2 \mathrm{\cos} \frac{\pi}{5} \simeq
1.6180$ is an algebraic integer that satisfies the equation:
\begin{equation}
\label{equation for k}
k^2 - k - 1 = 0.
\end{equation}
$\mathcal{G}$ has two linearly independent eigenvectors with
eigenvalue 0, which may be chosen as:
\[ n_{\left( 1 \right)} \equiv \left( \begin{array}{ccccccccc}
- 3 k - 2, & - 4 k - 2, & - 3 k - 1, & - 2 k, & - k, & k + 1, & k + 1, &
1, & 0
\end{array} \right) \]
\begin{equation}
\label{eigvecs with eigval 0} n_{\left( 2 \right)} \equiv \left(
\begin{array}{ccccccccc}
- 5 k - 3, & - 6 k - 4, & - 4 k - 3, & - 2 k - 2, & - 1, & k + 1, & k + 1,
& 0, & 1
\end{array} \right)
\end{equation}
If we regard $\mathcal{G}_{i j}$ as a ``metric'' in 9 dimensions, then for any
contravariant 9-vector $v^i$ such that $v^i v_i \neq 0$, where $v_i \equiv v^j
\mathcal{G}_{j i}$, the matrix $r \left( v \right)^i \, \!_j \equiv \delta^i
\, \!_j - 2 \frac{v^i v_j}{v^k v_k}$ satisfies $r \left( v \right)^i \, \!_k r
\left( v \right)^k \, \!_j = \delta^i \, \!_j$ and $\mathcal{G}_{k l} r \left(
v \right)^k \, \!_i r \left( v \right)^l \, \!_j =\mathcal{G}_{i j}$, and is
thus a reflection matrix with respect to the ``metric'' $\mathcal{G}_{i j}$. \
In particular, if we define $u_{\left( i \right)}^j$, $1 \leq i \leq 9$, to be
the vector whose $i \,$th component is 1 and whose other components are 0,
then since $u_{\left( i \right)}^j u_{\left( i \right) j} = 2$ for all $1 \leq
i \leq 9$, and all the matrix elements of $\mathcal{G}$ are in the ring of
algebraic integers $\mathbf{Z} \left[ k \right]$, the matrices $r \left(
u_{\left( i \right)} \right)^j \, \!_k$, $1 \leq i \leq 9$, are 9 reflection
matrices whose matrix elements are in $\mathbf{Z} \left[ k \right]$. \ We
now define $\tilde{u}^j_{\left( i \right)} \equiv u^j_{\left( i \right)}$, $1
\leq i \leq 7$, $\tilde{u}^j_{\left( 8 \right)} \equiv u^j_{\left( 8 \right)}
- n^j_{\left( 1 \right)}$, and $\tilde{u}^j_{\left( 9 \right)} \equiv
u^j_{\left( 9 \right)} - n^j_{\left( 2 \right)}$. \ Then $\tilde{u}_{\left( i
\right)}^j \tilde{u}_{\left( i \right) j} = 2$ for all $1 \leq i \leq 9$, and
the last two components of $\tilde{u}_{\left( i \right)}^j$ are 0 for all $1
\leq i \leq 9$. \ Let indices $a, b, c, \ldots$ run from 1 to 7,
$\mathcal{H}_{a b}$ be the $7 \times 7$ symmetric matrix $\mathcal{G}_{a b}$,
and $\bar{u}^a_{\left( i \right)}$, $1 \leq i \leq 9$, be the 7-vectors
$\tilde{u}_{\left( i \right)}^a$. \ Then for all $1 \leq i \leq 9$, the
matrices $r \left( \bar{u}_{\left( i \right)} \right)^a \, \!_b$ are $7 \times
7$ reflection matrices with respect to the ``metric'' $\mathcal{H}_{a b}$, and
their matrix elements are in $\mathbf{Z} \left[ k \right]$. \ Let:
\begin{equation}
\label{transformation matrix T} S \equiv \left( \begin{array}{ccccccc}
1 & \frac{k}{2} & \frac{3 k + 1}{5} & \frac{2 k + 1}{2} & 5 k + 3 & 0 &
0\\
0 & 1 & \frac{2 k + 4}{5} & k + 1 & 6 k + 4 & 0 & 0\\
0 & 0 & 1 & \frac{k + 2}{2} & 4 k + 3 & 0 & 0\\
0 & 0 & 0 & 1 & 2 k + 2 & 0 & 0\\
0 & 0 & 0 & 0 & 1 & 0 & 0\\
0 & 0 & 0 & 0 & 0 & 1 & \frac{k}{2}\\
0 & 0 & 0 & 0 & 0 & 0 & 1
\end{array} \right)
\end{equation}
Then $S^T \mathcal{H}S = \mathrm{{diag}} \left( \begin{array}{ccccccc}
2, & \frac{3 - k}{2}, & \frac{6 - 2 k}{5}, & \frac{2 - k}{2}, & - 2 k, & 2,
& \frac{3 - k}{2}
\end{array} \right)$, so $\mathcal{H}$ has signature $\left( 6, 1 \right)$,
and $\mathcal{H}_{a b}$ is the metric on 7-dimensional Minkowski space in a
non-standard coordinate system. \ The dot products $\bar{u}^a_{\left( i
\right)} \bar{u}_{\left( j \right) a} =\mathcal{H}_{a b} \bar{u}^a_{\left( i
\right)} \bar{u}^b_{\left( j \right)}$ are equal to the corresponding elements
of $\mathcal{G}_{i j}$ for all $1 \leq i, j \leq 9$, so the matrices $r \left(
\bar{u}_{\left( i \right)} \right)^a \, \!_b$ provide a representation of the
Coxeter group $\Gamma \left( T_3 \right)$ corresponding to the Coxeter diagram
$T_3$. \ Let superscript $^g$ denote the replacement of $k$ by its Galois
conjugate $k^g = \frac{1 - \sqrt{5}}{2} \simeq - 0.6180$, which is the other
solution of (\ref{equation for k}). \ Then the diagonal matrix elements of
$S^{g T} \mathcal{H}^g S^g$ are $> 0$, so $\mathcal{H}^g$ has signature
$\left( 7, 0 \right)$, so $\Gamma \left( T_3 \right)$ is of the arithmetic
type {\cite{Borel Harish Chandra}} reviewed in {\cite{Witte Morris}} and
subsection 3.1 of {\cite{CCHT}}.
A finite index torsionless normal subgroup $\Lambda$ of $\Gamma \left( T_3
\right)$ can be obtained by the same variant of Selberg's lemma
{\cite{Selberg}} as used in {\cite{Everitt Maclachlan}} to obtain the
4-dimensional Davis manifold from a compact hyperbolic Coxeter
4-simplex. \ $\Lambda$
consists of the elements of $\Gamma \left( T_3 \right)$ whose matrices in the
representation just constructed are equal $\textrm{{mod}}\, \sqrt{5}$ to
the unit matrix. \ To calculate the index of $\Lambda$ in $\Gamma$, which
gives the ratio of the intrinsic volume of the corresponding compact
hyperbolic 6-manifold to the intrinsic volume of the compact hyperbolic
Coxeter polytope $\xi \left( T_3 \right)$, we note that $\frac{1 +
\sqrt{5}}{2} + \sqrt{5} \frac{\sqrt{5} - 1}{2} = 3$, hence $k = 3\,
\textrm{{mod}}\, \sqrt{5}$. \ Thus the factor group $\Gamma / \Lambda$,
which is the group of all the distinct elements of $\Gamma\, \textrm{{mod}}\,
\sqrt{5}$, is obtained by replacing $k$ by 3 in the $7 \times 7$ matrix
representations of the 9 generators of $\Gamma$ constructed above, then
reducing the resulting integer matrix elements $\textrm{mod 5}$ and
doing the matrix multiplications with $\textrm{mod 5}$ arithmetic.
Using GAP {\cite{GAP}} we find $\mathrm{{Size}} \left( \Gamma / \Lambda
\right) = \text{460687500000000} = 2^8 3^4 5^{12} 7 \, 13 \simeq 4.6 \times
10^{14}$. \ Using the same procedure for the 6-dimensional compact hyperbolic
Coxeter polytope with 9 5-dimensional faces whose Coxeter diagram is shown in
Figure 4.3 of {\cite{Tumarkin}}, whose Gram matrix times 2 has matrix elements
in $\mathbf{Z} \left[ k \right]$ after dividing 2 rows and 2 columns by
$\sqrt{2}$, and which is also of arithmetic type, we find that in that case,
$\mathrm{{Size}} \left( \Gamma / \Lambda \right) = \text{914004000000000}
= 2^{11} 3^4 5^9 7 \, 13 \, \, 31 \simeq 9.1 \times 10^{14}$. \ The third
6-dimensional compact hyperbolic Coxeter polytope with 9 5-dimensional faces
is obtained from the Coxeter polytope $\xi \left( T_3 \right)$ by cutting it
in half along its hyperplane of reflection symmetry. \ By {\cite{Belolipetsky
2}} and Table 2 of {\cite{Belolipetsky 1}} the magnitudes of the fractional
Euler numbers of these three 6-dimensional compact hyperbolic Coxeter
polytopes are not less than $\frac{67}{1152000} \simeq 5.8 \times 10^{- 5}$,
so the intrinsic volumes of the compact hyperbolic 6-manifolds constructed in
this way are too large by a factor of not less than around $10^7$. \ However
there are other methods of constructing torsionless subgroups of hyperbolic
Coxeter groups {\cite{Everitt Maclachlan, Everitt, Conder Maclachlan, Long}}
that might provide examples of smaller intrinsic volume.
While version 3 of this article was being prepared, the XENON100
collaboration published the results of the first 100.9 days of data taking
in the XENON100 direct search for dark matter, finding no evidence for dark
matter interacting with the ultra-low background innermost 48 kg of their
62 kg liquid xenon target, and excluding spin-independent elastic
WIMP-nucleon scattering cross-sections above $7.0\times 10^{-45}$ cm$^2$ for
a WIMP mass of 50 GeV at 90\% confidence level \cite{XENON100}. The model
considered in this article could provide a variety of dark matter candidates
with masses up to around a TeV that interact only gravitationally with
ordinary matter, and are thus consistent with these limits, if there were
from 1 to around 10 other HW boundaries distributed around the compact
hyperbolic 7-manifold $\bar{H}^7$, possibly with different $E_8$ breakings
on each of them.
\begin{center}
{\bf Acknowledgements}
\end{center}
\noindent I would like to thank the organizers of the 2007 CERN BSM Institute,
in particular Nima Arkani-Hamed, Savas Dimopolous, and Christophe Grojean, for
arranging for me to give a talk and spend a very enjoyable and useful week at
CERN with financial support, Asimina Arvanitaki, Savas Dimopoulos, Philip
Schuster, Jesse Thaler, Natalia Toro, and Jay Wacker for very interesting
discussions, Greg Moore for a helpful email about flux quantization, Kasper
Peeters for correspondence about Cadabra both on and off the mailing list, and
Peter Woit \cite{Peter Woits blog} and ``Jester'' \cite{Jesters blog} for
providing efficient digests of what's happening
in high energy physics. \ The calculations made heavy use of Maxima
{\cite{Maxima}} and Cadabra \cite{Cadabra 1} and also used GAP \cite{GAP},
the diagram was drawn with help from TexPict \cite{TexPict}, the bibliography
was sequenced with help from Ordercite \cite{Ordercite},
and the article and the notebooks the work was done in were written with
GNU TeXmacs {\cite{TeXmacs}}
running in KDE 3.5 {\cite{KDE}} in Debian GNU/Linux {\cite{Debian}}.
\vspace{0.5cm}
|
\section{Introduction}
Pulsar timing observations \citep[for a review of the techniques,
see][]{lk05,ehm06} provide a unique opportunity to study
low-frequency ($10^{-9}$--$10^{-7}$ Hz) gravitational waves
\citep[GWs;][]{saz78,det79,bcr83,hd83,fb90,ktr94,jhl+05}. Previous
work \citep{rt83,ktr94,lom02,jhv+06} placed upper limits on a
stochastic background of GWs. These limits were reported in terms of
either the amplitude of the GW characteristic strain spectrum,
$h_c(f)$, or the normalized GW energy density, $\Omega_{\rm gw}(f)$.
In recent years, researchers have proposed that supermassive
black-hole (SMBH) binary systems distributed throughout the universe
will be a source of GWs detectable using pulsar timing techniques
\citep[e.g.,][]{jb03,wl03,eins04,shm+04,svc08}. The detection, or
non-detection, of such GWs provides a constraint on the rate of
coalescence of SMBH binary systems. We emphasize that such constraints
are model-independent as opposed to the indirect constraints that can
be inferred from observed galaxy distributions.
We will show that existing pulsar data sets do not provide stringent
constraints on the coalescence rate. However, future data sets from
the Parkes Pulsar Timing Array \citep[PPTA;][]{man08,hbb+09} and
similar North American \citep[NANOGrav;][]{jen09} and European timing
array \citep[EPTA;][]{skl+06} projects aim to produce data sets on 20 or
more pulsars with root-mean-square (rms) timing residuals close to
100\,ns. In the longer term, we expect that pulsar timing array projects using
future telescopes such as the Square Kilometre Array
(SKA)\footnote{See www.skatelescope.org} will be able to time many
hundreds of pulsars with exquisite timing precision.
The outline of this paper is as follows. In \S~2, the physics of GW
emission from SMBH binary systems is reviewed together with the
effects of GWs on pulsar timing. We describe how to constrain the
coalescence rate using two different techniques, one valid when there
is a large number of expected sources and the other valid when only a
few sources are expected. In \S~3 we show the recent and projected
rate constraints for various different observing systems. These
observationally constrained rates are then compared to the rates
implied by local galaxy-merger observations. This work is summarized
in section \S~4.
\section{Constraining the SMBH Merger Rate}
We define a SMBH as a black hole with mass greater than
$10^6~\Msolar$. There is abundant evidence that such SMBHs exist both
nearby \citep[e.g.,][]{kr92,mmh+95} and at high redshifts $z \sim$ 6
\citep[e.g.,][]{fnl+01}. Two orbiting SMBHs resulting from a galaxy
merger would emit large amplitude
GWs. Such GW sources are important targets for space-based detectors
such as the Laser Interferometer Space Antenna (LISA) and pulsar
timing arrays \citep[e.g.,][]{wl03,eins04,shm+04}. Observational GW
astronomy can be used to resolve whether SMBH binary systems form and
whether the two black holes can get close enough to emit detectable
GWs.
This work focuses on the rate constraints measurable by pulsar timing
observations. In order to determine the coalescence rate, one needs to
know the expected GW amplitude emitted by a SMBH binary. This is given
by \citep{tho87}
\begin{equation}
h_s=4\sqrt{\frac{2}{5}}\frac{(GM_c)^{5/3}}{c^4D(z)}[\pi f(1+z)]^{2/3},
\label{amp}
\end{equation}
where $M_c$ is the ``chirp mass'' of the SMBH binary given by $M_c
=(M_1M_2)^{3/5}(M_1+M_2)^{-1/5}$, $M_1$ and $M_2$ are the individual
black hole masses, $f$ is observed GW frequency, $D(z)$ is the
comoving distance to the system
\begin{equation}
D(z)=\frac{c}{H_0}\int_0^z
\frac{dz'}{E(z')},
\label{distance}
\end{equation}
with $E(z)=\sqrt{\Omega_{\Lambda}+\Omega_m(1+z)^3}$ for a $\Lambda$CDM
cosmological model (hereafter we adopt $H_0=72~{\rm km~s^{-1}~Mpc^{-1}}$,
$\Omega_m=0.3$, $\Omega_{\Lambda}=0.7$). GWs from such a source will induce
sinusoidally oscillating arrival-time variations whose amplitude,
$\Delta t$, is given by
\citep[][and references therein]{jll+04}:
\begin{equation}
\Delta t = \frac{h_s}{\omega} [1+\cos(\theta)]\sin(2\phi)\sin\{\omega D[1 - \cos(\theta)]/2c\},
\end{equation}
where $\omega$ is the GW frequency in radians/s, $\theta$ is the angle
on the sky between the pulsar direction and the GW source direction,
$\phi$ is the GW polarization angle, and $D$ is the distance to the
pulsar. The maximum induced timing residual amplitude, $h_s/\omega$,
is plotted in Figure \ref{rvsz} as a function of redshift for chirp
masses of $10^9$ and $10^{10}~\Msolar$ and observed GW frequencies of
(1 year)$^{-1}$ and (10 year)$^{-1}$. The most notable feature in this
Figure is that these amplitude curves are not monotonically decreasing
with increasing redshift. The reason for this is that the observing
frequency is held fixed and the frequency in the frame of the
emitting system increases with increasing redshift. For a binary
system, the emitted GW amplitude increases with increasing
frequency. For large $z$, this increase is faster than the decrease
due to increasing comoving distance $D(z)$. Also note that the curves
cutoff at large $z$. This is because there is a maximum orbital
frequency allowed before the black holes plunge together. This maximum
frequency was taken to be $c^3/(6^{3/2}\pi GM)$ assuming a circular
orbit \citep{hug02}. Here, $M=M_1+M_2$ is the total mass of a binary
system.
Rms timing residuals (for a typical 1-hour observation) for the best
pulsar data sets are currently around 100\,ns. Multiple observations
combined with improved systems will bring the effective sensitivity down
to around 10\,ns. In this case, Figure \ref{rvsz} shows that pulsar
timing will be sensitive to individual SMBH binary systems with chirp
masses greater than about $10^9~\Msolar$. Figure \ref{rvsz} also shows
that this sensitivity extends to large redshifts. This fact greatly
increases the chances of detecting individual sources. An ensemble of
lower-mass SMBH binary systems will be detectable as a stochastic
background if there is a large enough population of sources.
\begin{figure}
\epsscale{0.5}
\plotone{fig1.ps}
\caption{\label{rvsz}Induced timing residual amplitudes versus
redshift of a system with a given observed frequency and chirp mass.}
\vspace{2.em}
\end{figure}
Since pulsar timing techniques are sensitive to SMBH binary systems up
to high redshifts, the non-detection of any sources, either individual
binaries or a background generated by an ensemble of binary systems,
may be used to place a direct constraint on $d^2R/dM_c dz$, the
sky-averaged rate of coalescence of binary SMBHs per unit chirp mass,
$M_c$, per unit redshift $z$. The total number of binary SMBHs with
chirp mass between $M_c$ and $M_c + \Delta M_c$ and located between
$z$ and $z+\Delta z$ merging between time $t$ and $t + \Delta t$ is
given by $\Delta M_c \Delta z \Delta t d^2R/dM_cdz$. Constraints
placed on this quantity may be used to rule out various binary SMBH
formation models.
We present two methods for determining the coalescence rate from
pulsar timing data. The first is valid when the rate is high enough
that the GWs form a stochastic background (the ``stochastic
constraint'') and there is a large number of sources per resolvable
frequency bin. The second method (the ``Poissonian constraint'')
provides an estimate of the coalescence rate when the stochastic
constraint does not hold. For a real data set it is practical first
to assume the stochastic constraint, determine the coalescence rate
and check whether the rate is high enough for the assumption to be
valid. If not then the Poissonian constraint should be used.
\subsection{Stochastic constraint}
Here, it is assumed that a large number of SMBH binary sources form an
incoherent background of GWs. The power spectrum of such a background
is given by \citet{jb03}
\begin{equation}
P(f) = \int_{0}^{\infty} \int_{0}^{\infty} h_s(f,M_c,z)^2 \frac{d^2R}{dz dM_c} \left(\frac{df}{dt}\right)^{-1} dz dM_c,
\label{power1}
\end{equation}
where $h_s(f,M_c,z)$ is given by Equation~\ref{amp} and $df/dt$ is the
rate of change of the observed GW frequency. For the case of a binary
system evolving under general relativity alone, this is given by \citep{pm63}
\begin{equation}
\frac{df}{dt} = \frac{96}{5}\left(\frac{G M_c}{c^3}\right)^{5/3}\left(\pi f\right)^{8/3}f(1+z)^{5/3}.
\label{dfdt}
\end{equation}
Typically, a stochastic background of GWs is described by its
characteristic strain spectrum, $h_c(f)$, which is assumed to take on
a power law form:
\begin{equation}\label{eqn:bkgrd}
h_c(f)=A\Big(\frac{f}{f_{\rm yr}}\Big)^{\alpha},
\end{equation}
where $f_{\rm yr} = 1/(1 {\rm year})$ and $A$ is the characteristic
strain at a period of one year. For a background generated by SMBH
binaries, $\alpha=-2/3$ in frequency range
$10^{-9}$--$10^{-7}$ Hz \citep{jb03,wl03,shm+04,eins04}. Note that the
characteristic strain spectrum is related to the power spectrum by
$P(f) = h_c(f)^2/f$.
Pulsar timing data sets provide an upper bound, $A_{\rm up}$, on
$A$. Such bounds limit the power spectrum of the GW strain. The upper
bound, $P_{\rm up}(f)$, may be written as
\begin{equation}
P_{\rm up}(f) = \frac{A^2_{\rm up}}{f_{\rm yr}}\Big(\frac{f}{f_{\rm yr}}\Big)^{-7/3}.
\label{Pup}
\end{equation}
Since this is an upper bound, we have
\begin{equation}
P(f) < P_{\rm up}(f).
\label{ub}
\end{equation}
In order to use Equations~\ref{power1} and \ref{ub} to obtain a
constraint on the differential rate of coalescence itself, the
integrand is rewritten in an equivalent form:
\begin{equation}
P(f) = \int_0^{\infty} \int_{-\infty}^{\infty} h_s(f,M_c,z)^2 \frac{d^2R}{d\lg(1+z) d\lg(M_c)} \left(\frac{df}{dt}\right)^{-1} d\lg(1+z) d\lg(M_c).
\label{power2}
\end{equation}
Note that both $d^2R/d\lg(1+z) d\lg(M_c)$ and $df/dt$ depend on $z$ and
$M_c$, although the explicit dependence is not written. A constraint
on $P(f)$ is a direct constraint on the integral in the above
expression. In order to obtain an estimate of the upper bound on the
integrand, we follow the same line of reasoning used to place
constraints on the differential energy density of gravitational waves
using bounds from big bang nucleosynthesis \citep{mag00}. First, we
note that the limits in the integral of equation \ref{power2} are from
0 to $\infty$ for $\lg(1+z)$ and from $-\infty$ to $\infty$ for
$\lg(M_c)$. Consider this integral over a small region bounded
by $\lg(M_{c_1})$ to $\lg(M_{c_2})$ and $\lg(1+z_1)$ to $\lg(1+z_2)$.
Denote this integral as $P_s(f)$ where
\begin{equation}
P_s(f) = \int_{\lg(M_{c_1})}^{\lg(M_{c_2})} \int_{\lg(1+z_1)}^{\lg(1+z_2)} h_s(f,M_c,z)^2
\frac{d^2R}{d\lg(1+z) d\lg(M_c)} \left(\frac{df}{dt}\right)^{-1} d\lg(1+z) d\lg(M_c).
\end{equation}
The mean value theorem tells us that there exist values of $M_c$ and
$z$, written as $M_c^*$ and $z^*$, such that
\begin{equation}
P_s(f) = h_s(f,M_c^*,z^*)^2 \frac{d^2R}{d\lg(1+z) d\lg(M_c)}
\left(\frac{df}{dt}\right)^{-1} \Delta\lg(1+z) \Delta\lg(M_c),
\end{equation}
where $\Delta\lg(1+z) = \lg(1+z_2) - \lg(1+z_1)$ and $\Delta\lg(M_c) =
\lg(M_{c_2}) - \lg(M_{c_1})$. Next, we assume that the integrand
varies slowly over the region of integration so that $P_s(f)$ does
not change much as long as $M_c^*$ and $z^*$ are chosen within this
region. From this, we see that $P_s(f)$ is
approximately given by
\begin{equation}
P_s(f) \approx h_s(f,M_c,z)^2 \frac{d^2R}{d\lg(1+z) d\lg(M_c)}
\left(\frac{df}{dt}\right)^{-1} \Delta\lg(1+z)\Delta\lg(M_c)
\label{power_approx}
\end{equation}
for any value of $M_c \in [M_{c_1}, M_{c_2}]$ and $z \in
[z_1,z_2]$. Next, since the integrand in equation \ref{power2} is
positive definite, it follows that
\begin{equation}
P_s(f) \leq P(f).
\label{ub2}
\end{equation}
From equations \ref{Pup}, \ref{ub}, \ref{power_approx}, and \ref{ub2} we have:
\begin{equation}
h_s(f,M_c,z)^2 \frac{d^2R}{d\lg(1+z) d\lg(M_c)} \left(\frac{df}{dt}\right)^{-1} \Delta\lg(1+z)\Delta\lg(M_c)
\lesssim \frac{A^2_{\rm up}}{f_{\rm yr}}\Big(\frac{f}{f_{\rm yr}}\Big)^{-7/3},
\label{integrand_c}
\end{equation}
where we are free to choose any value for $M_c$ and $z$ provided that
the integrand is slowly varying over the appropriate region. Using the
known expressions for $h_s(f,M_c,z)$ and $df/dt$, one can use equation
\ref{integrand_c} to obtain a constraint on the
differential rate of coalescence. Assuming $\Delta\lg(M_c)=1$ and
$\Delta \lg(1+z) = 0.2$, the constraint takes the form
\begin{equation}
\frac{d^2 R}{d\lg(1+z) d\lg(M_c)} < 15 A^2_{\rm up} \frac{c^3 D(z)^2 (1 + z)^{1/3}}{(G M_c)^{5/3}}
(\pi f_{\rm yr})^{4/3},
\label{const_st}
\end{equation}
which is independent of frequency. Therefore, a measured bound,
$A_{\rm up}$, can be used directly to constrain the SMBH coalescence
rate. Note that, if one believes that the integrand changes more rapidly with $z$
and/or $M_c$, one can choose a sufficiently small integration range
over which this assumption is true and then rescale the above
constraint.
This constraint is only valid when there are a large number of sources
emitting into the same frequency band at the same time. In order for
this to be true, the GW amplitude of each source must be much less
than the minimum detectable amplitude as determined by the statistical
properties of the pulsar timing data, otherwise a detection would
have been made. This reasoning leads to the following constraint:
$h_s(f,M_c,z)^2 \ll P_{\rm up}(f) \Delta f$, where $\Delta f$ is the
resolution bandwidth which is taken to be $1/T_{\rm obs}$. For
the purposes of making numerical estimates, the following constraint is
used
\begin{equation}
h_s(f,M_c,z)^2 \leq 0.1 \frac{P_{\rm up}(f)}{T_{\rm obs}}.
\end{equation}
For a fixed chirp mass and frequency, this expression is
a constraint on $z$. The most stringent constraint occurs when
$f=1/T_{\rm obs}$, the lowest observable frequency. Combining the above
with Equations~\ref{amp} and \ref{Pup} yields the following constraint
on the redshift
\begin{equation}
\frac{(1+z)^{2/3}}{D(z)} \leq 0.19\Big(\frac{A_{\rm up}}{10^{-14}}\Big)
\Big(\frac{T_{\rm obs}}{\rm yr}\Big)^{4/3}
\Big(\frac{M_c}{10^8~\Msolar}\Big)^{-5/3}{\rm Mpc}^{-1}.
\label{condition}
\end{equation}
The factor $(1+z)^{2/3}/D(z)$ decreases with $z$ until $z=$2.65,
where it starts to increase. Hence, there is a bounded redshift
interval over which the stochastic constraint is valid. For systems
outside of this range, the stochastic rate limit is not valid and the
Poisson rate limit discussed in the next section must be employed.
\subsection{Poisson constraint}
For this case, the sources are not numerous enough to form a
stochastic background, so they must be treated as individual
events. Assuming Poissonian statistics for the probability of an event
occurring, the probability that no events are detected is given by
$e^{-\langle N\rangle}$, where $\langle N\rangle$ is the expected
number of events. Since no events are detected in the pulsar timing
data, the upper limit on the expected number, $\langle N_*\rangle$, is
set so that $e^{-\langle N_*\rangle} = 0.05$. Hence, $\langle N\rangle
\le \langle N_*\rangle = 3$. If the actual expected number were
greater than $3$, then at least one source would have been detected
with 95\% probability.
Provided that the expected number of events that occur within the
resolution bandwidth is less than one, the expected number of
detectable events is given by
\begin{equation}
\langle N \rangle = \int\frac{d^2R}{d\lg(1+z) d\lg(M_c)}\left(\frac{df}{dt}\right)^{-1}P_d(M_c,z,f) d\lg(1+z) d\lg(M_c)df,
\label{exp_num}
\end{equation}
where $P_d(M_c,z,f)$ is the probability of detecting a SMBH binary
with chirp mass $M_c$ at a redshift of $z$ with observable frequency
$f$. $P_d$ takes into account non-GW noise sources that can reduce the
GW detection efficiency. Following the same argument used in the
previous section to obtain an upper bound on the differential rate
using an integral constraint, we find that:
\begin{equation}
\frac{d^2R}{d\lg(1+z)d\lg(M_c)}< \frac{15}{\int \left(\frac{df}{dt}\right)^{-1}P_d(M_c,z,f) df}.
\label{const_pn}
\end{equation}
As with the stochastic constraint, it is assumed that
$\Delta\lg(M_c)=1$ and $\Delta \lg(1+z) = 0.2$. This constraint should
be appropriately rescaled if the integrand in equation \ref{exp_num}
varies over shorter intervals.
The above constraint requires a knowledge of $P_d$, the probability of
detecting a SMBH binary system using pulsar timing data. The $P_d$
is calculated using a method similar to that in \citet{yhj+10}.
For our analysis we use a Monte-Carlo simulation together with the
data analysis pipeline \citep{hjl+09} used to search for GWs in real and
simulated pulsar timing data. We use a Neyman-Pearson decision
technique together with a Lomb periodogram to determine the
probability of detection. For a set of $N_p$ pulsars, the power
spectra of the timing residuals are calculated and added
together. This summed power spectrum is used as the detector. We
determine noise levels and hence detection thresholds for each
frequency channel so that the false alarm rate for a detection is
0.001 across the entire power spectrum. The detection thresholds are
found by producing 100,000 fake data sets by shuffling the input
timing residuals, carrying out standard pulsar timing fits and forming
the summed power spectrum.
Once the thresholds are determined, the probability of detecting a GW
with a given strain amplitude is calculated as follows. A GW strain
amplitude $A$ and frequency $f$ are chosen (this procedure is repeated
on a logarithmically spaced grid where $A$ ranges from
$10^{-16}$--$10^{-10}$ and $f$ ranges from 1/(30 years)--1/(2
weeks)). The GW polarization properties are chosen to be consistent
with GWs emitted from a binary system and simulated as described in
\citet{hjl+09}. The direction of the GW wave-vector is chosen from a
distribution that is uniform on the sky while its polarization is
drawn from a distribution of randomly oriented binary systems. The
induced timing residuals from this GW source are added to a shuffled
version of the original residuals and the summed periodogram is
calculated. This is repeated 1000 times and $P_d(A,f)$ is given by the
number of times that the GW was detected (i.e. produced power above
the threshold) divided by the total number of trials.
\section{Results and Discussion}
\begin{deluxetable}{lccc}
\tablecolumns{4}
\tablewidth{0pc}
\tablecaption{Upper limits on the amplitude of the stochastic GW
background. A '--' indicates that there is no valid range for that scenario.
\label{range}}
\tablehead{
\colhead{Data set} & {$A_{\rm up}$} & \multicolumn{2}{c}{Valid redshift range} \\
\cline{3-4}
& & {($10^9~\Msolar$)}& {($10^{10}~\Msolar$)}
}
\startdata
PPTA data \citep{jhv+06}&$1.1\times10^{-14}$& 0.01--180.08& 1.78--4.02 \\
20 PSRs-500\,ns-10\,yr & $1.1\times10^{-15}$& 0.03--602.60& -- \\
20 PSRs-100\,ns-5\,yr & $9.9\times10^{-16}$& 0.07--294.97& -- \\
20 PSRs-100\,ns-10\,yr & $2.2\times10^{-16}$& 0.14--115.98& -- \\
20 PSRs-10\,ns-10\,yr & $2.0\times10^{-17}$& -- & -- \\
100 PSRs-100\,ns-5\,yr & $5.7\times10^{-16}$& 0.14--121.46& -- \\
100 PSRs-100\,ns-10\,yr& $1.3\times10^{-16}$& 0.27--45.30 & -- \\
100 PSRs-10\,ns-10\,yr & $8.8\times10^{-18}$& -- & -- \\
\enddata
\end{deluxetable}
The expressions in the previous sections can be used to provide
constraints on the rate of coalescence with any measurement of
$A_{\rm up}$ (for the stochastic case) or $P_d$ (for the Poissonian case).
Here we discuss the implications of the value of $A_{\rm up}$ presented by
\cite{jhv+06}, use the same data set to determine $P_d$ and simulate
data sets predicting possible future timing residuals. For these
future data sets we simulate (1) a realistic goal for
existing pulsar timing array experiments (20 pulsars,
timed with an rms timing residual of 500\,ns over 10 years), (2) the
goal of the PPTA project (20 pulsars, timed with
an rms of 100\,ns over 5 years), and a more challenging goal of
20 pulsars, timed with an rms of 100\,ns over 10 years. The rms
timing residuals that will be achieved with future telescopes, such as
the SKA, is not easy to determine. It may be possible to time a few
pulsars with exquisite precision (with rms timing residuals of 10\,ns)
but other unmodeled noise processes may make this difficult or impossible. We
therefore also simulate the following possible future data sets: (4)
100 pulsars timed at 100\,ns over five years, (5) the same for ten
years, (6) 20 pulsars timed at 10\,ns for 10 years and (7) the same
for 100 pulsars.
Figures ~\ref{rate_st1} and \ref{rate_st2} plot the stochastic
constraints given by equation~\ref{const_st} for the different
observing scenarios. The horizontal error bars indicate the region of
$\lg(1+z)$ over which the constraint is placed. The solid error bars
indicate that the stochastic constraint is valid, while the dotted
error bars indicate that the stochastic validity condition is violated
over the whole range. Table~\ref{range} gives the valid redshift range
for each data set and chirp mass. As discussed above, it was assumed
that $\Delta\lg(1+z)=0.2$ and $\Delta\lg(M_c)=1$ in order to calculate
the constraints shown. The constraints should be rescaled if other
values are assumed.
For redshifts where the stochastic technique is not valid, the
Poissonian constraint may be used. Using $P_d(A,f)$,
Equation~\ref{amp} is used to calculate $P_d(M_c,z,f)$ and then the
right hand side of Equation~\ref{const_pn} is evaluated numerically to
determine the constraint on the differential coalescence rate. The
results are plotted versus redshift for different chirp masses in
Figures~\ref{rate_pn1} and \ref{rate_pn2}. We note that the recently published
PPTA data set, that with 20 pulsars timed with an rms of 500\,ns over
10\,years, and that with 20 pulsars timed with an rms of 100 \,ns over
5\,years are not constraining for SMBH binaries with
$M_c=10^9~\Msolar$ and are therefore not plotted in the left-hand
panel of Figure~\ref{rate_pn1}.
In order to understand how constraining the pulsar rate limits are,
plots of the expected SMBH binary coalescence rate are also shown in
the figures. Several authors have developed analytical and numerical
techniques to estimate the coalescence rate of SMBH binaries. Here, we
compare the measured rate constraints to the rates predicted by the
models of \citet{jb03} and \citet{svc08,svv09}. For the case of
\citet{jb03}, the differential rate of SMBH binary coalescence is
given by
\begin{equation}
\frac{d^2R}{d\lg(1+z) d\lg(M_c)}=\frac{4\pi c}{H_0}\frac{D(z)^2}{\lg(e)E(z)}
\frac{\Phi(\lg (M_c))}{n}\Re(z),
\label{ratezm}
\end{equation}
where $\Re(z)$ is the total merger rate of SMBH binaries per unit comoving volume,
$\Phi(\lg(M_c))$ is the chirp mass distribution of merging SMBH binaries and $n$ is
the number density of SMBH binaries given by $n=\int \Phi(\lg(M_c))d\lg (M_c)$. It is
assumed that the merger rate of SMBH binaries is given by a fraction,
$\epsilon$, of the galaxy merger rate and that the rate evolves as a
power of $(1+z)$. Hence, one can write $\Re(z) = \epsilon \Re_g(0)(1
+ z)^\gamma$ where $\Re_g(0)$ is the local merger rate of galaxy pairs
and $\gamma$ is the evolution index which is thought to be
within the range $-1 < \gamma < 3$
\citep[e.g.,][]{ppc+02,lkw+04,kss+07,lpk+08}. \citet{wlh09} determined
$\Re_g(0)$ and $\Phi(\lg(M_c))$ for luminous galaxies by analyzing data
from the Sloan Digital Sky Survey \citep[SDSS;][]{yaa+00}. They found
that $\Re_g(0) = (1.0\pm0.4)\times 10^{-5}~{\rm ~Mpc^{-3}~Gyr^{-1}}$ and
\begin{equation}
\lg [\Phi(\lg (M_c))]=(21.7\pm4.2)-(3.0\pm0.5)\lg \left(M_c/\Msolar\right).
\label{chmass}
\end{equation}
Note that \citet{wlh09} showed that the rate implied by equation
\ref{ratezm} together with the above estimates for $\Re_g(0)$ and
$\Phi(\lg (M_c))$ yield an expected characteristic strain spectrum
consistent with other published estimates of $h_c$
\citep{wl03,shm+04,eins04,svc08}.
For the \citet{svc08,svv09} models, the rate
may be estimated from the following expression:
\begin{equation}
\frac{d^2R}{d\lg(1+z) d\lg(M_c)}= \frac{d\dot{N}_m}{d\lg(M_c)}
\frac{1}{\dot{N}}\frac{d\dot{N}}{dz}\frac{(1+z)}{\lg(e)}
\label{rate_sesana}
\end{equation}
where $d\dot{N}_m/d\lg(M_c)$ is the mass function of coalescing SMBHs in
the notation of \citet{svv09}, and $d \dot{N}/dz$ is the SMBH binary
coalescence rate per unit redshift in the notation of
\citet{svc08}. The constant $\dot{N}$ is given by
\begin{equation}
\dot{N} = \int_0^4 \frac{d\dot{N}}{dz} dz.
\end{equation}
The limits of the integral are set by the data presented in figure 12
of \citet{svc08}. The factor of $(1+z)/\lg(e)$ is used to convert the
differential $dz$ into $d\lg(1+z)$. From figure 1 of \citet{svv09}, we
can estimate the maximum and minimum predicted values of the mass
function $d\dot{N}_m/d\lg(M_c)$ over the four models presented
therein. We find that, for $M_c=10^9 \Msolar$, the mass function lies
between $10^{-4} ~\mbox{yr}^{-1}$ and $6\times 10^{-3}
~\mbox{yr}^{-1}$. These correspond to the ``Tr-SA'' and the ``La-SA''
models, respectively, as discussed in \citet{svv09}. For the case of
$M_c=10^{10} \Msolar$, the predicted range lies between $0$ an $3
\times 10^{-5} ~\mbox{yr}^{-1}$. These values also correspond to the
``Tr-SA'' and ``La-SA'' models, respectively. For the SMBH binary coalescence rate
per unit redshift, $d\dot{N}/dz$, we used the data presented in
Figure 12 of \citet{svc08}. The three possible models shown are all
approximately within a factor of two of each other. For definiteness,
we chose the prediction based on the BVRhf model. In this case,
$\dot{N}~\approx 0.05 ~\mbox{yr}^{-1}$.
Figures \ref{rate_st1} and \ref{rate_st2} plot the pulsar timing
stochastic constraint together with the expected rates for the
theoretical models considered above as a function of redshift for
different chirp masses. The Poissonian constraint is plotted together
with the expected rates in Figures \ref{rate_pn1} and
\ref{rate_pn2}. Since there are few to no close SMBH binary systems
detected near $z=0$, it can be assumed that $\epsilon \approx 1$. The
only free parameter remaining in the \citet{jb03} model is the
evolution index which determines the SMBH binary merger rate as a
function of redshift. The grey regions in Figures
\ref{rate_st1}--\ref{rate_pn2} give the range of expected merger rates
for $-1 < \gamma < 3$. The maximum and minimum rates found using the
\citet{svc08,svv09} models are shown as thick dashed and thin
dashed lines, respectively. Note, since the minimium predicted rate for
$M_c=10^{10} \Msolar$ is not presented by the Sesana et al. models, this curve is
not shown. Overall, the upper bounds obtained by pulsar timing data do
not contrain the parameters of the SMBH binary merger models discussed in
this paper beyond their currently accepted ranges. For the PPTA goal (20 pulsars timed at
100\,ns rms accuracy for five years), the results imply that either a
detection will be made or $\gamma<1.7$ at redshift $z<3$. In order to
place constraints that will limit the models of \citet{svc08,svv09} as
well as the \citet{jb03} model with $\gamma < -1$, one must either
time 100 pulsars with 100\,ns rms timing precision or 20 pulsars at
the 10\,ns level, both of which should be possible with the proposed
Square Kilometre Array project.
The Poissonian constraint is only useful for constraining the
properties of the most massive SMBH binaries since these systems are
rarer than their less massive counterparts and emit a stronger GW
signal. Figure \ref{rate_pn1} shows that the ideal PPTA extended to 10
years of observations will just be able to place useful limits for the
most massive systems if $\gamma$ were in the larger end of its
possible range. It will be more interesting when we can time pulsars
to the 10\,ns level. Here, the Poissonian constraint will be well
below that expected for $\sim$$10^9~\Msolar$ SMBH binaries from both
the \citet{svc08,svv09} and \citet{wlh09} models. Hence, with 20
pulsars timed with 10\,ns rms timing accuracy for 10 years, we have a
very good chance of detecting an individual source or will place very
stringent constraints on models of SMBH binary formation and
evolution. These conclusions are consistent with the recent work of
\citet{svv09}.
\begin{figure*}
\epsscale{0.9}
\plotone{fig2.ps}
\caption{Upper limits on the SMBH binary merger rate determined by the
stochastic constraint discussed in \S2.1 for different data sets: real
data published by \citet{jhv+06} (open triangle), simulated data for
20 PSRs-500\,ns-10\,yr (open square), 20 PSRs-100\,ns-5\,yr (cross),
and 20 PSRs-100\,ns-10\,yr (open circle). The error bar is plotted as
a dotted line when the constraint is invalid. The filled gray area
represents the expected region for the coalescence rate using the
framework of \citet{jb03} together with the data from the SDSS
\citep{wlh09} with an evolution index $-1 < \gamma < 3$. The dashed
lines between $0<\lg(1+z)<0.7$ indicate the maximum (thick) and
minimum (thin) predicted rates from \citet{svc08,svv09}.
\label{rate_st1}}
\end{figure*}
\begin{figure*}
\epsscale{0.9}
\plotone{fig3.ps}
\caption{Same as Figure~\ref{rate_st1}, but for simulated data sets:
100 PSRs-100\,ns-5\,yr (filled triangle), 100 PSRs-100\,ns-10\,yr
(filled square), 20 PSRs-10\,ns-10\,yr (star), and 100 PSRs-10\,ns-10\,yr
(filled circle).
\label{rate_st2}}
\end{figure*}
\begin{figure*}
\epsscale{0.9}
\plotone{fig4.ps}
\caption{Upper limits on the SMBH merger rate using the Poissonian
constraint discussed in \S2.2 for different data sets: real data from \citet{jhv+06}
(open triangle), 20 PSRs-500\,ns-10\,yr (open square), 20
PSRs-100\,ns-5\,yr (cross), and 20 PSRs-100\,ns-10\,yr (open
circle). The filled gray area represents the expected region for
the coalescence rate using the framework of \citet{jb03} together
with the data from the SDSS \citep{wlh09} with an evolution index,
$-1 < \gamma < 3$. The dashed lines between $0<\lg(1+z)<0.7$
indicate the maximum (thick) and minimum (thin) predicted rates from
\citet{svc08,svv09}. Note that no upper limits on the coalescence
rate are obtainable for SMBH
of $M_c=10^9~\Msolar$ at these sensitivities.
\label{rate_pn1}}
\end{figure*}
\begin{figure*}
\epsscale{0.9}
\plotone{fig5.ps}
\caption{Same as Figure~\ref{rate_pn1}, but for the following
simulated data sets: 100 PSRs-100\,ns-5\,yr (filled triangle), 100 PSRs-100\,ns-10\,yr
(filled square), 20 PSRs-10\,ns-10\,yr (star), and 100 PSRs-10\,ns-10\,yr
(filled circle).
\label{rate_pn2}}
\end{figure*}
\section{Summary}
\label{discuss}
We have shown that pulsar timing observations may be used to place
constraints on the rate of coalescence of binary supermassive black
holes distributed throughout the Universe. Two types of constraints
were considered: a stochastic constraint and a Poissonian
constraint. The stochastic constraint, which is based on a detection
algorithm for the stochastic GW background, gives lower rates but it
is only valid when the expected amplitude for a individual source is
much less than the minimum detectable amplitude. When this is not the
case, the Poissonian constraint must be used. This constraint is based
on a continuous-wave detection algorithm and it assumes that the
number of coalescence events are distributed according to a Poisson
distribution. In both cases, it is assumed that the differential rate
of coalescence varies sufficiently slowly over a range of chirp masses
and redshifts. The precise numerical value of the constraint depends
on the size of the interval over which the rate is assumed to be
nearly constant.
The implied rate constraint obtained from recently published data
together with rate constraints expected from future possible observing
scenarios were compared to theoretical rates calculated from different
models. It was shown that 20 pulsars timed with an accuracy of
100\,ns, the goal of the PPTA project, will place stringent
constraints on the semi-empirical models based on the work of
\citet{jb03} and \citet{wlh09}. The upper end of the range of
backgrounds produced by SMBH population synthesis models
\citep{svc08,svv09} would be detectable by the PPTA goal sensitivity,
but higher sensitivity will be needed to further constrain these models
if the GW background is not detected. It was also shown that if future
observations can time a pulsar with 10\,ns accuracy, a direct
detection of one or more individual sources is highly likely.
\acknowledgments
We thank Professor Han J. L. for his support and valuable comments. FJ
was supported by the National Science Foundation (CAREER grant no. AST
0545837). GH is the recipient of an Australian Research Council QEII
Fellowship (project \#DP0878388). ZW is supported by the National
Natural Science Foundation of China (projects \#10521001 and
\#10833003). FJ also thanks these projects for local support during
his visit at the NAOC for two months in 2006. The authors also wish to
thank the anonymous referee for all her/his useful suggestions and
comments during the review process. The Parkes telescope is part of
the Australia Telescope which is funded by the Commonwealth Government
for operation as a National Facility managed by CSIRO.
|
\section{Introduction}
The large-scale structure of the universe can be relatively well-explained by cold dark matter (CDM) models. In these models, the dark matter is represented by a collisionless gas without pressure (dust) interacting via Newtonian gravity only. It can be described by the Vlasov-Poisson system or, under additional approximations, by the pressureless Euler-Poisson system in an expanding background \cite{peeblesbook}. However, this model faces several problems on the scale of galactic or sub-galactic structures. For example, CDM simulations lead to dark matter halos with density cusps \cite{nfw} while observations of rotation curves favor flat density profiles. In addition, it predicts an abundance of satellite galaxies around each galactic halo that is far beyond what we see around the Milky Way. At the cosmological level, nonlinear gravitational clustering is often studied by resorting to the Zeldovich approximation \cite{zeldovich}. This approximation leads to the inviscid Burgers equation which describes the free motion of fluid particles \cite{vergassola}. However, this equation develops shocks and formal singularities known as caustics (having the form of pancakes), so that it becomes invalid after particle-crossing. If we accept multi-stream solutions, the particles just cross each other after the caustic and pass through the pancakes instead of clustering into smaller objects like groups of galaxies. Therefore, a small-scale regularization must be introduced to overcome this problem and model the effects of ``punctuated'' gravitational attraction and pressure gradients that are not captured by the Zeldovich approximation. Gurbatov {\it et al.} \cite{gurbatov} introduced the so-called ``adhesion model'' in which particles move according to the Zeldovich approximation until they fall into pancakes when their trajectories intersect, then ``stick'' to each other. This sticking can be modeled by introducing a viscosity in the Burgers equation in order to represent the effect of strong gravitational forces or pressure effects in the vicinity of a caustic. Of course, the viscosity must be small in order to provide a smoothing effect at small-scales only (where particle-crossing occurs) but the limit $\nu\rightarrow 0$ is different from taking $\nu=0$. This model gives very good results in the nonlinear regime and can reproduce the skeleton of the ``cosmic web'' of large-scale structures (sheets, filaments, nodes) in $N$-body numerical simulations (see, e.g., Ref. \cite{ssmpm}). A completely different approach was developed by Widrow \& Kaiser \cite{wk} who proposed to describe a classical collisionless self-gravitating gas by the Schr\"odinger-Poisson system. In this approach, the constant $\hbar$ is not the Planck constant, but rather an adjustable parameter that controls the spatial resolution $\lambda_{deB}$ through a de Broglie relation $\lambda_{deB}=\hbar/mv$. It is argued that when $\hbar\rightarrow 0$, the Vlasov-Poisson system is recovered and that a finite value of $\hbar$ provides a small-scale regularization of the dynamics. In that case, the Schr\"odinger-Poisson system has nothing to do with quantum mechanics since it aims at describing the evolution of classical collisionless matter under the influence of gravity (in static or expanding universes). This model was further developed by Short \& Coles \cite{sc} who introduced a {\it free-particle approximation}. They showed that the dynamics of the particles before crossing is relatively close to the Zeldovich approximation but when crossing occurs the quantum pressure prevents the formation of singularities. Therefore, the quantum pressure replaces the role of the viscosity in the adhesion model. Although these models have given interesting results and can be very useful in practice, one can argue, however, that their justification remains relatively {\it ad hoc}.
Recently, B\"ohmer \& Harko \cite{bohmer} have proposed that dark matter could be a self-gravitating Bose-Einstein condensate (BEC) with quartic self-interaction described by the Gross-Pitaevskii-Poisson (GPP) system. This idea takes its origin in the concept of boson stars introduced by Kaup \cite{kaup} and Ruffini \& Bonazzola \cite{rb}. It is well-known that the Gross-Pitaevskii equation (or the nonlinear Schr\"odinger equation) can be reduced to hydrodynamical equations by means of the Madelung \cite{madelung} transformation. This yields the Euler-Poisson system with a ``classical'' isotropic pressure due to self-interaction and a quantum anisotropic pressure arising from the Heisenberg uncertainty principle. For a quartic self-interaction, the barotropic equation of state is that of a polytrope of index $n=1$. At large-scales, the pressure terms are negligible and one recovers the CDM model which has proven very successful. However, at small-scales, the classical and quantum pressures can play a crucial role and regularize the dynamics by preventing the formation of singularities (caustics). BEC dark matter has very interesting properties: (i) the pressure can stabilize the system against gravitational collapse and lead to dark matter halos with flat density profiles. For very light bosons without self-interaction, the stabilization is due to the quantum pressure. For more massive bosons with self-interaction, the stabilization is due to repulsive scattering. Therefore, a BEC dark matter can solve the cusp problem and the missing satellite problem \cite{hu,bohmer,paper1}. (ii) BEC dark matter halos can reproduce the rotation curves of several low surface brightness galaxies \cite{arbey,bohmer}. (iii) At a cosmological level, a BEC dark matter can accelerate the formation of structures with respect to ordinary CDM models \cite{harko,cosmobec}. In this brief report, we make the additional remark that if dark matter is a self-gravitating BEC with quartic self-interaction, the adhesion approximation, the Burgers equation and the cosmological KPZ equation that have been introduced phenomenologically to solve the problems inherent to CDM models find a natural justification and an interesting generalization.
\section{Stochastic Gross-Pitaevskii-Poisson system}
\label{sec_gpp}
We assume that dark matter is a self-gravitating BEC with quartic self-interaction described by the stochastic Gross-Pitaevskii-Poisson (GPP) system
\begin{equation}
\label{gpp1}
i\hbar \frac{\partial\psi}{\partial t}=-\frac{\hbar^2}{2m}\Delta\psi+m(\Phi+g Nm|\psi|^2+\eta)\psi,
\end{equation}
\begin{equation}
\label{gpp2}
\Delta\Phi=4\pi G Nm |\psi|^2,
\end{equation}
where $\psi({\bf r},t)$ is the wave function, $\rho({\bf r},t)=Nm|\psi|^2$ the density, $\Phi({\bf r},t)$ the gravitational
potential, $g={4\pi a_s\hbar^2}/{m^3}$ the pseudo-potential accounting for short-range interactions ($a_s$ is the s-scattering length) \cite{revuebec} and $\eta({\bf r},t)$ a stochastic potential (noise). We write the wave function in the form $\psi({\bf
r},t)=A({\bf r},t)e^{iS({\bf r},t)/\hbar}$ where $A$ and $S$ are real,
and make the Madelung \cite{madelung} transformation $\rho=Nm|\psi|^2=NmA^2$ and ${\bf u}=\nabla S/m$,
where $\rho({\bf r},t)$ is the density field and ${\bf u}({\bf r},t)$ the velocity field. We note that the flow is irrotational since $\nabla\times {\bf u}={\bf 0}$. With this transformation, the stochastic GP equation (\ref{gpp1}) is {\it equivalent} to the stochastic barotropic Euler equations with an additional term $Q=-(\hbar^2/2m)\Delta\sqrt{\rho}/\sqrt{\rho}$ called the quantum potential (or quantum pressure). Indeed, one obtains the set of equations
\begin{equation}
\label{gpp4}
\frac{\partial\rho}{\partial t}+\nabla\cdot (\rho {\bf u})=0,\qquad \Delta\Phi=4\pi G \rho,
\end{equation}
\begin{equation}
\label{gpp5}
\frac{\partial {\bf u}}{\partial t}+({\bf u}\cdot \nabla){\bf u}=-g\nabla \rho-\nabla\Phi+\frac{\hbar^2}{2m^2}\nabla\left (\frac{\Delta \sqrt{\rho}}{\sqrt{\rho}}\right )-\nabla\eta.
\end{equation}
The first term on the r.h.s. of the Euler equation (\ref{gpp5}) can be interpreted as a classical isotropic pressure $-(1/\rho)\nabla p$ described by a barotropic equation of state $p=p(\rho)$. For a quartic self-interaction, it is given by
\begin{equation}
\label{gpp7}
p=\frac{1}{2}g\rho^2=\frac{2\pi a_s\hbar^2}{m^3}\rho^{2}.
\end{equation}
This is a polytropic equation of state of the form $p=K\rho^{\gamma}$
with polytropic index $n=1$ (i.e. $\gamma=1+1/n=2$) and polytropic constant $K=g/2={2\pi a_s\hbar^2}/{m^3}$. Other types of barotropic equations of state can be obtained depending on the form of the self-interaction. A detailed study of the time-independent solutions of the GPP system (or quantum barotropic Euler-Poisson system), connecting the non-interacting limit ($a_s\simeq 0$) \cite{rb} to the Thomas-Fermi (TF) limit ($Q\simeq 0$) \cite{bohmer} has been given recently in Ref. \cite{paper1}.
\section{BEC equations in an expanding universe}
\label{sec_exp}
For the sake of simplicity, we consider an expanding Einstein-de Sitter (EdS) universe ($\Lambda=0$ and $\kappa=0$) \cite{peeblesbook} described by the equations
\begin{equation}
\label{exp1}
\rho_b a^3\sim 1,\qquad \ddot a=-\frac{4}{3}\pi G \rho_b a,\qquad {\dot a}^2=\frac{8}{3}\pi G\rho_b a^2,
\end{equation}
yielding $a\propto t^{2/3}$, $H=\dot a/{a}={2}/{3t}$ and $\rho_b={1}/{6\pi Gt^2}$, where $\rho_b(t)$ is the background density, $a(t)$ the scale factor and $H={\dot a}/a$ the Hubble constant. Our approach can be easily extended to more general models of universe.
We shall first rewrite the hydrodynamic equations in the comoving frame. Making the change of variables ${\bf r}=a(t){\bf x}$, ${\bf u}=H{\bf r}+{\bf v}$ and $\phi=\Phi-\Phi_b$ where ${\bf v}$ is the peculiar velocity and $\Phi_b=(2/3)\pi G\rho_b(t)r^2$ the background gravitational potential, and introducing the density contrast $\delta({\bf x},t)=\lbrack \rho({\bf x},t)-\rho_b(t)\rbrack/\rho_b(t)$, we obtain the system of equations \cite{cosmobec}:
\begin{equation}
\label{exp3}
\frac{\partial\delta}{\partial t}+\frac{1}{a}\nabla\cdot \lbrack (1+\delta) {\bf v}\rbrack=0,\qquad \Delta\phi=4\pi G \rho_b a^2\delta,
\end{equation}
\begin{eqnarray}
\label{exp4}
\frac{\partial {\bf v}}{\partial t}+\frac{1}{a}({\bf v}\cdot \nabla){\bf v}+\frac{\dot a}{a}{\bf v}=-\frac{g \rho_b}{a}\nabla \delta\nonumber\\
-\frac{1}{a}\nabla\phi+\frac{\hbar^2}{2m^2a^3}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right )-\nabla\eta.
\end{eqnarray}
For simplicity, we shall always denote the noise by $\eta$ although a new notation should be introduced after each transformation. At large scales, pressure and noise effects are negligible and the CDM model ($\hbar=p=\eta=0$) is recovered. However, pressure effects become important when nonlinear structures form, and the BEC dark matter model (\ref{exp3})-(\ref{exp4}) should be used instead.
Measuring the evolution in terms of $a$ rather than in terms of $t$ and introducing the new velocity ${\bf w}={\bf v}/a\dot a$ and the new gravitational potential $\psi=\phi/4\pi G\rho_b a^3$, we can recast the foregoing equations in the form
\begin{equation}
\label{exp6}
\frac{\partial\delta}{\partial a}+\nabla\cdot \lbrack (1+\delta) {\bf w}\rbrack=0, \qquad \Delta\psi=\frac{\delta}{a},
\end{equation}
\begin{eqnarray}
\label{exp7}
\frac{\partial {\bf w}}{\partial a}+({\bf w}\cdot \nabla){\bf w}+\frac{3}{2a}({\bf w}+\nabla\psi)=-\frac{3 a_s\hbar^2}{2Gm^3a^4}\nabla \delta\nonumber\\
+\frac{3\hbar^2}{16\pi G m^2\rho_b a^6}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right )-\nabla\eta.
\end{eqnarray}
For short times, the perturbations are small $\delta\ll 1$, $\phi\ll 1$, $|{\bf w}|\ll 1$, and the hydrodynamic equations can be linearized. We obtain
\begin{equation}
\label{lin1}
\frac{\partial\delta}{\partial a}+\nabla\cdot {\bf w}=0,\qquad \Delta\psi=\frac{\delta}{a},
\end{equation}
\begin{eqnarray}
\label{lin2}
\frac{\partial {\bf w}}{\partial a}+\frac{3}{2a}({\bf w}+\nabla\psi)=-\frac{3 a_s\hbar^2}{2Gm^3a^4}\nabla \delta\nonumber\\
+\frac{3\hbar^2}{32\pi G m^2\rho_b a^6}\nabla(\Delta \delta)-\nabla\eta.
\end{eqnarray}
Taking the ``time'' derivative of Eq. (\ref{lin1}-a) and the divergence of Eq. (\ref{lin2}), these equations can be combined into a single equation for the density contrast
\begin{eqnarray}
\label{lin3}
\frac{\partial^2\delta}{\partial a^2}+\frac{3}{2a}\frac{\partial\delta}{\partial a}=\frac{3\delta}{2a^2}\nonumber\\
+\frac{3a_s\hbar^2}{2Gm^3a^4}\Delta\delta-\frac{3\hbar^2}{32\pi G m^2\rho_b a^6}\Delta^2\delta+\Delta\eta.
\end{eqnarray}
This equation has been studied in Ref. \cite{cosmobec} in the no-noise limit.
At large scales, we can ignore pressure and noise effects, and we recover the equation for the density contrast of a cold gas
\begin{eqnarray}
\label{lin4}
\frac{\partial^2\delta}{\partial a^2}+\frac{3}{2a}\frac{\partial\delta}{\partial a}=\frac{3\delta}{2a^2}.
\end{eqnarray}
The growing solution to this equation is $\delta_+({\bf x},a)=aD({\bf x})$ \cite{peeblesbook}. Then, Eq. (\ref{lin1}-b) implies that $\psi({\bf x},a)=\psi({\bf x})$ is constant. Therefore, $\delta_+({\bf x},a)=a\Delta\psi({\bf x})$. On the other hand, in the cold gas approximation, Eq. (\ref{lin2}) reduces to
\begin{eqnarray}
\label{lin5}
\frac{\partial {\bf w}}{\partial a}+\frac{3}{2a}({\bf w}+\nabla\psi)={\bf 0}.
\end{eqnarray}
After a transient regime, the velocity field tends toward the solution ${\bf w}({\bf x},a)=-\nabla\psi({\bf x})$.
\section{Zeldovich approximation and a generalized equation of structure formation}
\label{sec_z}
The Zeldovich approximation \cite{zeldovich} amounts to extending this relation to the (weakly) nonlinear regime, i.e. ${\bf w}({\bf x},a)\simeq -\nabla\psi({\bf x},a)$. The fact that ${\bf w}$ is a potential flow in the BEC model gives further support to this approximation. In that case, Eq. (\ref{exp7}) reduces to the form
\begin{eqnarray}
\label{z1}
\frac{\partial {\bf w}}{\partial a}+({\bf w}\cdot \nabla){\bf w}=-\frac{3 a_s\hbar^2}{2Gm^3a^4}\nabla \delta\nonumber\\
+\frac{3\hbar^2}{16\pi Gm^2\rho_b a^6}\nabla \left (\frac{\Delta \sqrt{1+\delta}}{\sqrt{1+\delta}}\right )-\nabla\eta,
\end{eqnarray}
where the explicit dependence on the gravitational potential has disappeared. Since $\delta=a\Delta\psi=-a\nabla\cdot {\bf w}$ and $\nabla\delta=-a\Delta {\bf w}$, we can rewrite the foregoing equation as
\begin{eqnarray}
\label{z2}
\frac{\partial {\bf w}}{\partial a}+({\bf w}\cdot \nabla){\bf w}=\frac{3 a_s\hbar^2}{2Gm^3a^3}\Delta{\bf w}
\nonumber\\
+\frac{3\hbar^2}{16\pi Gm^2\rho_b a^6}\nabla \left (\frac{\Delta \sqrt{1-a\nabla\cdot {\bf w}}}{\sqrt{1-a\nabla\cdot {\bf w}}}\right )-\nabla\eta.
\end{eqnarray}
This can be viewed as a generalized noisy Burgers equation with an additional quantum pressure term. Since ${\bf w}$ is a potential flow, we obtain
\begin{eqnarray}
\label{z3}
\frac{\partial \psi}{\partial a}=\frac{(\nabla\psi)^2}{2}+\frac{3 a_s\hbar^2}{2Gm^3a^3}\Delta\psi\nonumber\\
-\frac{3\hbar^2}{16\pi Gm^2\rho_b a^6} \frac{\Delta \sqrt{1+a\Delta\psi}}{\sqrt{1+a\Delta\psi}}+\eta.
\end{eqnarray}
This can be viewed as a cosmological KPZ equation \cite{kpz}. Without the forcing term, this is just the Bernouilli equation, or the Hamilton-Jacobi equation, with an additional quantum potential. We note that the potential $\psi$ is related to the phase $S$ of the wave function of the BEC by
$\psi=-(1/a)(S/ma^2H-x^2/2)$. A noisy Burgers equation and a KPZ equation have been introduced and studied previously in cosmology by Buchert {\it et al.} \cite{buchert}, Jones \cite{jones}, Coles \cite{coles} and Matarrese \& Mohayaee \cite{mm}.
Let us consider particular cases of Eq. (\ref{z2}).
(i) For $a_s=\hbar=\eta=0$, we get the inviscid Burgers equation which is equivalent to the Zeldovich approximation.
(ii) If we neglect the quantum potential, we get the noisy Burgers equation
\begin{eqnarray}
\label{z5}
\frac{\partial {\bf w}}{\partial a}+({\bf w}\cdot \nabla){\bf w}=\nu(a)\Delta{\bf w}-\nabla\eta,
\end{eqnarray}
with a time-dependent viscosity given by $\nu(a)={3 a_s\hbar^2}/{2Gm^3a^3}$ (this equation is time-reversible via $a\rightarrow -a$ and ${\bf w}\rightarrow -{\bf w}$).
Therefore, the BEC dark matter hypothesis leads to a natural justification of the (stochastic) adhesion model. Another justification was previously given by Buchert {\it et al.} \cite{buchert} in terms of a coarse-graining process inherent to a hydrodynamic description. However, to arrive at the Burgers equation, several simplifying hypothesis had to be introduced: (i) dark matter is described by hydrodynamic equations with an {\it isotropic} pressure; (ii) the equation of state is a polytrope $p=K\rho^{\gamma}$ with $\gamma=2$; (iii) the velocity field and the stochastic force derive from a potential. Interestingly, these properties directly result from the BEC dark matter hypothesis (i.e. from the GP equation) without further assumption. Note also that the change of variable $\psi({\bf x},a)=2\nu(a)\ln W({\bf x},a)$ transforms Eq. (\ref{z3}) into
\begin{eqnarray}
\label{z3b}
\frac{\partial W}{\partial a}=\nu(a)\Delta W+\left\lbrack \frac{1}{2\nu}\eta({\bf x},a)+\frac{3}{a}\ln W\right \rbrack W.
\end{eqnarray}
(iii) In the absence of self-interaction and noise, we obtain an equation of the form
\begin{eqnarray}
\label{z7}
\frac{\partial {\bf w}}{\partial a}+({\bf w}\cdot \nabla){\bf w}=\frac{3\hbar^2}{16\pi Gm^2\rho_b a^6}\nabla \left (\frac{\Delta \sqrt{1-a\nabla\cdot {\bf w}}}{\sqrt{1-a\nabla\cdot {\bf w}}}\right ).\nonumber\\
\end{eqnarray}
This can be viewed as a Burgers equation where the small-scale regularization is provided by the quantum pressure. Our approach has some similarities with the effective wave mechanics approach of Short \& Cole \cite{sc}. However, in our approach, the coefficient in front of the quantum potential in Eq. (\ref{z7}) depends on ``time'' $a$, while in the approach of Short \& Cole \cite{sc}, this coefficient is constant (see their Eq. (24)). This is because they introduce an effective Schr\"odinger equation [their Eq. (23)] directly in the comoving frame while we start from the (nonlinear) Schr\"odinger equation (\ref{gpp1}) in the inertial frame. When we write the (nonlinear) Schr\"odinger equation (\ref{gpp1}) in the comoving frame, we obtain an equation different from their Eq. (23). Ribeiro \& Peixoto de Faria \cite{rp} have also developed an effective wave mechanics approach in which they relate the gradient of the quantum potential to the Laplacian of a kinematical velocity plus a noise term. Their approach, which is essentially phenomenological, leads to a result different from Eq. (\ref{z7}).
\section{Conclusion}
\label{sec_conclusion}
In this brief report, we have shown that the assumption that dark matter is a self-gravitating Bose-Einstein condensate with quartic self-interaction described by the (stochastic) GPP system leads to a natural justification of the adhesion model, the Burgers equation and the cosmological KPZ equation without more hypothesis than the Zeldovich approximation that is common to most works on the subject. Therefore, not only the BEC model is consistent with previous works, but it generalizes them and extends their scope. In addition, it gives a new justification of these phenomenological models. This result adds to the other nice properties of BEC dark matter (flat density profiles, flat rotation curves, acceleration of the growth of perturbations...) recalled in the Introduction.
The BEC model not only follows the general evolution of inhomogeneities but it also describes the internal structure of density enhancements. In this sense, it improves upon the standard adhesion model. Indeed, at the level of dark matter halos, the equations reduce to the condition of hydrostatic equilibrium and lead to virialized structures similar to $n=1$ polytropes (other configurations could be obtained depending on the form of self-interaction and on the equation of state). Therefore, BEC dark matter provides a model that describes both the large-scale structures of the universe (through the generalized Burgers equation (\ref{z2})) and the structure of dark matter halos (through the explicit expression of the pressure (\ref{gpp7}) arising from short-range interactions).
If we justify the GPP system in terms of quantum mechanics, we must assume that the mass $m$ of the bosons is extraordinarily small ($m< 10^{-24}\, {\rm eV}/c^2$!) for quantum mechanics to be relevant on galactic scales. This has been called ``fuzzy dark matter'' in Ref. \cite{hu}. Alternatively, one can produce similar quantum effects with larger boson masses if the particles have a self-interaction \cite{colpi}. Since the nature of dark matter is not known, we cannot reject {\it a priori} that quantum mechanics plays some role at galactic or cosmological scales. In any case, the GPP system (\ref{gpp1})-(\ref{gpp2}) can always be viewed as an effective wave mechanics approach with an adjustable Planck constant which generalizes the initial model of Widrow \& Kaiser \cite{wk} based on the Schr\"odinger-Poisson (SP) system. Finally, other approaches to the problem, based on the Vlasov-Poisson (VP) system, could be contemplated, see e.g. Ref. \cite{csr}.
|
\section{Introduction}
Discovering a few years ago a new family of iron based superconductors gave rise both to intensive investigation of the fundamental properties of these materials (see reviews \cite{Sad,Ivan,Izy,Ishida}
and Refs. therein) and to seeking for their potential application.
The fabrication of high-quality thin films is of great importance for potential applications of these new materials in
superconducting devices as well as for a deeper fundamental study of their superconducting properties.
Numerous experiments were undertaken to study the superconducting state of iron based superconductors, however less
attention was paid to the investigation of films, due to the more complicated preparation.
In this paper we present first point-contact Andreev reflection (PCAR) spectroscopy investigation of
LaFeAsO$_{1-x}$F$_{x}$ and SmFeAsO$_{1-x}$F$_{x}$ films.
The main goal was to measure of the superconducting gap(s) and its temperature dependence for the mentioned
films to compare these data with existing similar measurements on bulk samples \cite{Chen,Gonnelli,Wang}
and to clarify some issues as to the presence of multiband structure on PCAR spectra and PCAR spectra asymmetry.
\section{Experimental details}
$Re$FeAsO$_{1-x}$F$_{x}$ ($Re$=La, Sm) films with superconductivity onset below 34\,K have
been fabricated using pulsed laser deposition. The successful growth of high quality $Re$FeAsO$_{1-x}$F$_{x}$
thin films \cite{Backen,Kidszun,Kidszun1,Kidszun2} enables the investigation of fundamental properties.
Details of the preparation of LaFeAsO$_{1-x}$F$_{x}$ films are described in Ref. \cite{Kidszun}.
The same parameters were also used for SmFeAsO$_{1-x}$F$_{x}$ film growth.
The temperature dependence of the resistance of the $Re$=Sm film used in our experiments is
shown in Fig.\,\ref{rt}.
The onset of the superconducting transition of about 34\,K is well below the maximal transition temperature
for optimally doped SmFeAsO$_{1-x}$F$_{x}$ crystals likely due to severe fluorine loss under film preparation.
Also the superconducting transition sufficiently broadens in a magnetic field, what can be due to nonuniform
fluorine doping through the film thickness. The $Re$=La film was of lower quality with a much broader
superconducting transition. Therefore, we concentrated in this study mainly on measurements of the $Re$=Sm film.
\begin{figure} [b]
\vspace{3cm}
\begin{center}
\includegraphics[width=8.5cm,angle=0]{sfa1.eps}
\vspace{-3cm}
\end{center}
\vspace{0cm} \caption[] {Superconducting transition
of a SmFeAsO$_{1-x}$F$_{x}$ thin film at zero field and at 9\,T. Arrow shows $T_c^{onset}\simeq 34\,K$.
Inset: Temperature dependence of the resistance of the film up to room temperatures.} \label{rt}
\end{figure}
The PCs were established by the standard "needle-anvil" method \cite{Naid} touching of
the film surface by a sharpened thin Cu wire. The differential resistance $dV/dI(V)$
was recorded by sweeping the dc current $I$ on which a small ac current $i$ was superimposed
using the standard lock-in technique. The measurements of PCAR spectra were performed in the temperature range
between 3 and 33\,K applying a magnetic field up to 7\,T perpendicular to the film surface.
\section{Characteristic lengths}
In PCAR spectroscopy some conditions must be fulfilled, namely, the contact diameter $d$ has to be smaller
than the inelastic electron mean-free path $l_i$ as well as smaller than the superconducting coherence length $\xi$.
The latter is anisotropic in the $Re$FeAsO$_{1-x}$F$_{x}$ family and does not exceed a few nm \cite{Ishida}.
The contact size can be estimated from its resistance $R_{\rm {PC}}$ above $T_c$ using modified Wexler formula \cite{Naid}
for the case of heterocontact with a metal, where its resistivity is negligible compared with that of $Re$FeAsO$_{1-x}$F$_{x}$:
\begin{equation}
R_{\rm {PC}}\simeq (1+Z^2)\frac{16\rho l}{3 \pi d^2}+ \frac{\rho}{2d}.
\label{Wexler}
\end{equation}
Here, $Z$ is the so called barrier strength \cite{BTK}. We estimated the product $\rho l$ using the upper limit
for the electron carrier concentration $n_e\leq 10^{21} $cm$^{-3}$ as inferred from Hall data
\footnote{The Hall coefficient determined in \cite{Jaros} for $Re$=La is about 3 times lager
at low temperatures than that for $Re$=Sm, which points to a 3 times larger electron density in $Re$=Sm.
At the same time for $Re$=Sm with an onset of superconductivity at 36\,K, similar as
in our film, the electron density is estimated to be about $n_e\leq 0.6\cdot 10^{21} $cm$^{-3}$ \cite{Tropeano}.}
for $Re$=La just above $T_c$ \cite{Sefat}. Making use of the Drude free electron model we calculated
$\rho l \simeq 1.3\cdot10^{4} n_e^{-2/3} = 1.3\cdot 10^{-10}\Omega\cdot$cm$^{2}$.
The residual resistivity $\rho_0$ of $Re$FeAsO$_{1-x}$F$_{x}$ compounds, estimated by extrapolation $\rho(T\to0)$, has value of about 100\,$\mu\Omega\cdot$cm for $Re$=La, whereas $\rho_0$ is higher for other $Re$FeAsO$_{1-x}$F$_{x}$ compounds
\footnote{According to \cite{Jaros}, the resistivity of $Re$=Sm is 2-3 times larger at low temperatures
than that of $Re$=La.}. Using the calculated value of $\rho l$ and the mentioned resistivity $\rho_0$, we find 10\,nm
as upper limit for the elastic electron mean free path $l_e$ for $Re$FeAsO$_{1-x}$F$_{x}$ compounds.
In Table I, we present estimations of the PC diameter, according to Eq.\,(\ref{Wexler}), for PCs with three different
resistances using both the lower limit of $\rho_0$=100 $\mu\Omega\cdot$cm (for high quality samples) and the higher
value $\rho_0$=1\,m$\Omega\cdot$cm, taken from the literature for $Re$FeAsO$_{1-x}$F$_{x}$.
\begin{table}[b]
\caption{Calculation of the PC diameter for three PC resistances according to Eq.\,(\ref{Wexler}) for $Z$=0 and two values
of resistivity and for $Z$=0.6 (last column).}
\begin{tabular}{|c|c|c|c|}
\hline
$R_{\rm {PC}}$ & $\rho_0$=100 $\mu\Omega\cdot$cm & $\rho_0$=1 m$\Omega\cdot$cm & $\rho_0$=100 $\mu\Omega\cdot$cm, $Z$=0.6 \\
\hline
10$\Omega$ & 78\,nm & 504\,nm &85\,nm \\
30$\Omega$ & 37\,nm & 171\,nm &41\,nm \\
100$\Omega$& 18\,nm & 54\,nm &20\,nm \\
\hline
\end{tabular}
\end {table}
Thus, in order to fulfil the condition that the dimension of the PC has to be smaller than the elastic mean free path
of electrons and/or the coherence length the PC resistance must be well above a few hundred Ohms,
which is not the case for PCs both investigated in this paper and in existing
PCAR studies cited in Ref.\,\cite{Gonnelli}. Therefore all PCAR measurements have
been done at least in the diffusive regime when elastic mean free path is small, i. e. $d\gg l_e $. Although the diffusive regime
does not prevent PCAR spectroscopy, it favors the shortening of the diffusive inelastic
mean free path $\Lambda\simeq (l_e\cdot l_i)^{1/2}$ of electrons so that a transition to the thermal regime $d\gg $min$(l_i, \Lambda$) \cite{Naid,Verkin79} with increasing
of applied voltage is getting more probable. Taking into account that the resistivity in all $Re$FeAsO$_{1-x}$F$_{x}$ has a remarkable
slope (increase) just above $T_c$, e. g., for $Re$=La, $\rho(T)$ behaves like $T^2$ below $\sim$ 200\,K \cite{Sefat},
the inelastic, likely electron-electron, scattering in $Re$FeAsO$_{1-x}$F$_{x}$ is starting already at very low temperatures shortening an inelastic mean free path $ l_i$ and $\Lambda$.
\begin{figure} [t]
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa2.eps}
\end{center}
\vspace{0cm} \caption[] {(Color on-line) Upper panel: $dV/dI$ curves of a
SmFeAsO$_{1-x}$F$_{x}$ - Cu contact ($R$ = 26\,$\Omega$)
for varying temperature. Bottom panel: Temperature dependencies of the fitting
parameters: superconducting gap $\Delta$ (triangles), barrier parameter $Z$ (diamonds),
scaling parameter $S$ (squares), position of minima in $dV/dI$ (stars) for the PC
from the upper panel. The broadening parameter $\Gamma$ is equal
to zero. The solid line represents the BCS-like gap behavior.
Inset: symmetrized and normalized \footnote{Resistivity of Sm and La pnictides has pronounced temperature dependence even at low temperatures and also below $T_c$, as it follows from the measurement in a magnetic field, see, e.g., Ref.\,\cite{Kidszun2}. It means that, so-called, normal state curve used for normalization can be slightly different for different temperatures. Therefore we have used for normalization of each $dV/dI(V)$ a weak parabolic-like curve which fitted the same $dV/dI(V)$ at biases $|V|>$10\,mV, i.e. above the gap minima.} $dV/dI$ curve at
$T$ = 3\,K (points) together with curves calculated according to
the generalized BTK theory: dashed (blue) and solid (red) lines - fit with $\Delta$=3.35\,mV, $\Gamma$=0, $Z$=0.38, $S$=0.5 and
$\Delta$=3.47\,mV, $\Gamma$=0.25\,mV, $Z$=0.41, $S$=0.58, respectively.} \label{dvdi}
\end{figure}
On the other hand, the coherence length $\xi$ is also much smaller than the contact size,
what can result in a distribution of superconducting properties (e.\,g., critical temperature, gap value) within
the PC and in a suppression of superconductivity in some part of the PC with current increase.
Thereby, regardless of the PC resistance (of course the higher the resistance the higher
probability to be in the spectral regime), each $dV/dI(V)$ curve should be
critically analyzed if it suits for spectroscopy of the superconducting gap.
\section{PCAR spectroscopy of the superconducting energy gap in S{\tiny m}F{\tiny e}A{\tiny s}O$_{1-x}$F$_{x}$}
Because of mentioned above reasons, it was difficult to get "clean" PCAR $dV/dI$ spectra having no
humps, spikes and other irregularities, which are connected with deviations from the spectral
regime in PC. Fig.\,\ref{dvdi} shows one of the best series of $dV/dI$ curves,
which demonstrate clear Andreev-reflection (gap) structure, namely pronounced minima at $V\simeq \pm3$\,mV
at $T\ll T_{\rm c}$ which are to a great extend free from unwanted features.
To get the superconducting gap value $\Delta$ and other parameters from the PCAR spectra
the Blonder-Tinkham-Klapwijk (BTK) theory \cite{BTK} including so-called
broadening parameter $\Gamma$ has been used.
The temperature dependence of the superconducting gap $\Delta$ has been established (see Fig.\,\ref{dvdi})
from the fitting within the BTK theory. We tried to keep constant such fitting parameters as $\Gamma$,
the barrier $Z$ and the scaling $S$.
\begin{figure} [t]
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa3.eps}
\end{center}
\vspace{0cm} \caption[] {$dV/dI$ curves of several SmFeAsO$_{1-x}$F$_{x}$ -- Cu contacts measured at 3\,K.
PC resistance is shown for each curve. Vertical dashed stripes mark reproducible AR minima.} \label{dvdis}
\end{figure}
Here, the parameter $S$ corresponds to the ratio of the experimental $dV/dI$ intensity to the calculated one
\footnote{Parameter S is intended to equalize the intensities of experimental and theoretical curves.
Theoretically $S$ = 1, but $S <$ 1 happens very often too. The possible reasons are connected with deviation of the PC
structure from theoretical model and are described in more detail in Appendix of Ref.\,\cite{Bobrov}.}.
It is seen from Fig.\,\ref{dvdi} that $\Delta(T)$ is in line with the BCS curve (solid line),
however, the resulting critical temperature of $T_{\rm c}^{*}\simeq$ 21.6\,K obtained by extrapolating the BCS curve
to $\Delta$= 0 is significantly lower than $T_{\rm c}$ determined from the temperature dependence of the resistivity
of the SmFeAsO$_{1-x}$F$_{x}$ film (see Fig.\,\ref{rt}). At the same time, the superconducting main minimum in
$dV/dI$ for the investigated PC disappears close to 30\,K as shown in Fig.\,\ref{dvdi} which approximately
corresponds to the midpoint ($T_{\rm c}^{50\%}$) of the superconducting transition in Fig.\,\ref{rt}. However, by approaching
this temperature, the shape of the $dV/dI$ minimum starts to deviate from the theoretically expected behavior
already above 20\,K as shown in Fig.\,\ref{dvdi}.
This is likely due to inhomogeneities of the superconducting state in the PC region in the case if its size $d$ is larger
than the coherence length $\xi$. Also with increasing temperature inelastic scattering increases favoring the
transition to the thermal regime. This hinders a determination of the gap in this temperature range. Therefore, the
determination of the gap of the PC shown in Fig.\,\ref{dvdi} is restricted to temperatures below 20\,K.
Investigating a dozen of PCs with clear double minima AR structure around $\pm 3$\,mV (see, Fig.\,\ref{dvdis}), we,
nevertheless, could not observe or detect reproducible structures at a higher voltage, which were found, e.g., in the representative two-band
superconductor MgB$_2$ \cite{Naid1}. Also an about 3 times larger second gap features reported for bulk SmFeAsO$_{1-x}$F$_{x}$
samples \cite{Gonnelli} could not be unambiguously resolved for the investigated PCs of the SmFeAsO$_{1-x}$F$_{x}$ film. Instead,
for increasing bias usually a peaked structure or other irregularities appear in the $dV/dI$ characteristics testifying
the transition to the non-spectral regime. But even if the similar to the second gap features are observed as, e.g,
shallow minima slightly above $\pm 10$\,mV in the upper PC spectrum in Fig.\,\ref{dvdis} (see also low temperature spectra in Fig.\,\ref{dvdi}), then this gap structure expected around $\pm$(9-12)\,meV is hardly to resolve for the other PC spectra (see, e.g., the bottom spectrum in Fig.\,\ref{dvdis}).
Moreover, already a weak magnetic field suppresses the mentioned structure, nevertheless no presence of any second
gap features appears (see Fig.\,\ref{dvdim}). The same concerns $dV/dI$ of PC with $R =14\,\Omega$, where
a shoulder is seen above $\pm$ 10\,mV often taken as a larger gap structure. The shoulder is washed out in magnetic
field 7\,T, while the small-gap minima are only slightly suppressed at the same field. On the contrary, in the mentioned MgB$_2$
the small gap is vanished more quickly in magnetic field than the larger one (see circles in Fig.\,15 in \cite{YansonR}).
\begin{figure} [t]
\vspace{3cm}
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa2b.eps}
\vspace{-3cm}
\end{center}
\vspace{0cm} \caption[] {$dV/dI$ curves of PC from Fig.\,\ref{dvdi} measured at 3\,K in a magnetic field.
Inset shows dependence of $\Delta$ (triangles) and $\Gamma$ (squares) vs magnetic field as determined from the
BTK-fitting of $dV/dI$ curves from the main panel. Here, $Z\simeq$0.38 and $S\simeq$0.5. Linear extrapolation of
$\Delta$ to zero result in a critical field of about 30\,T.} \label{dvdim}
\end{figure}
The above mentioned larger gap structure is also badly resolved in a recent tunneling study of SmFeAsO$_{1-x}$F$_{x}$ compounds
\cite{Noat}. The authors concluded that interband quasiparticle scattering has a crucial effect on the shape of the
tunneling spectra smearing out in particular a larger gap structure in the electronic DOS. Taking into account the very short
elastic electron mean free path in SmFeAsO$_{1-x}$F$_{x}$ compounds the interband quasiparticle scattering is expected to be quite strong.
From another tunneling measurements \cite{Fasano}, a distribution of gaps (conductance peaks) $\Delta_p$ between 6 and 8\,meV
with a mean value of 7\,meV was derived, which results in a reduced gap 2$\Delta_p /kT_c\simeq3.6$. Also from these tunneling
spectra, no second gap structure was resolved.
We show in Fig.\,\ref{his} a histogram of the gap distribution measured for PCs with the pronounced double minimum AR structure.
The distribution has a maximum around $2\Delta$=6\,mV and a high energy tail around 10\,mV. We connect this
large gap value and its broad distribution with the inhomogeneities of the superconducting properties on the film
surface due to fluorine loss. Supposing that the critical temperature for the PC with the gap of
around 10\,meV is close to the onset of superconductivity of $\simeq$ 34\,K in Fig.\,\ref{rt},
then $2\Delta /kT_c$ will be again close to the BCS value of 3.5.
\begin{figure} [t]
\vspace{3cm}
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa5.eps}
\vspace{-3cm}
\end{center}
\vspace{0cm} \caption[] {Histogram of the gap distribution determined from $dV/dI(V)$ of SmFeAsO$_{1-x}$F$_{x}$ -- Cu contacts
with pronounced double minimum AR structure. Inset shows an example of $dV/dI(V)$ for PC with $2\Delta\simeq$10\,meV.}
\label{his}
\end{figure}
\section{Asymmetry of $dV/dI$ curves}
\begin{figure} [t]
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa4.eps}
\end{center}
\vspace{0cm} \caption[] {Left panels: $dV/dI$ curves of a SmFeAsO$_{1-x}$F$_{x}$ -- Cu contacts measured at 3\,K (solid curves)
for a broader bias range. Dashed curves have been measured for the same contacts at higher temperature. PC in panel (a) is
the same as shown in Fig.\,\ref{dvdi}. Right panels: antisymmetric part $dV/dI^{as}(\%)= 100R_{PC}^{-1}(dV/dI(V>0)-dV/dI(V<0))$
for the corresponding curves from the left panels. Voltage polarity is given for the film.} \label{as}
\end{figure}
A characteristic feature of $dV/dI(V)$ of PCs with $Re$FeAsO$_{1-x}$F$_{x}$ is their pronounced asymmetry. From our study of this phenomenon
we conclude that the asymmetry of $dV/dI(V)$ is present independent of the superconducting features. Fig.\,\ref{as}
shows $dV/dI(V)$ curves for a few PCs, where clear AR double minima are observed (Fig.\,\ref{as}a) or the superconducting
structure is superimposed on a broad maximum (Fig.\,\ref{as}c) or no superconducting features, but only a sharp zero-bias
maximum is present (Fig.\,\ref{as}e).
In all cases, the antisymmetric part of $dV/dI(V)$ shows similar behavior and even the same relative value independent
of the drastic difference in $dV/dI(V)$ shape corresponding to different mechanisms of $dV/dI(V)$ formation. Such robustness
of the asymmetry of $dV/dI(V)$ testifies that this phenomenon is not related to the PC properties, but mainly determined by
the properties of the bulk material (or the film). One common and plausible reason of $dV/dI(V)$ asymmetry can be the thermoelectric
(Seebeck) effect if the PC is heated up at increasing bias in the thermal regime \cite{Naid}. It is known that $Re$FeAsO$_{1-x}$F$_{x}$ compounds
have a huge Seebeck coefficient at low temperatures which reaches -80\,$\mu$V$\cdot$K$^{-1}$ for $Re$=Sm just above
$T_c$ for a sample with $T_c$=53\,K \cite{Matus} and even more (-140\,$\mu$V$\cdot$K$^{-1}$) for a sample with $T_c$=36\,K \cite{Tropeano}
(similar as in our film). Taking thermoelectric effects in the case of heterocontacts into account, the $dV/dI(V)$ asymmetry
was explained both for heavy-fermion high-resistive compounds \cite{Naid,Naid2} and for diluted Kondo-alloys \cite{Naid,Naid3},
where the Seebeck coefficient is strongly enhanced.
It seems that an equilibrium heating of the PC is not the only condition for the appearance of thermoelectric effects. The electron
distribution function in ballistic and diffusive PCs is in a highly nonequilibrium state which can be represented by two Fermi
spheres shifted from each other by applied voltage e$V$ \cite{Naid}. This nonequilibrium electronic state produces nonequilibrium
phonons with an effective temperature $\leq$eV/4 \cite{Kulik} which is close to the temperature eV/3.63 \cite{Verkin79} reached in
the thermal regime.
Therefore, we suggest that the nonequilibrium state in the PC can also cause a thermoelectric voltage in the case of heterocontacts.
This results in a $dV/dI(V)$ asymmetry similar as observed in PCs in the thermal regime \cite{Naid}.
We also would like to draw attention that asymmetry of the STM spectra, see Figs.\,5, 6 in Ref.\,\cite{Noat}, variates and even has different sign for the spectra with similar structure. Since STM spectra reflects in a direct way electronic density of states (DOS), then the non-reproducible asymmetry of STM spectra point out also to the non-DOS nature of the asymmetry, what strengthen our
arguments.
\section{PCAR spectroscopy of the superconducting energy gap in L{\tiny a}F{\tiny e}A{\tiny s}O$_{1-x}$F$_{x}$}
As mentioned above, the investigated LaFeAsO$_{1-x}$F$_{x}$ film has a rather poor quality showing a broad superconducting transition already at zero-field which is comparable with that of the SmFeAsO$_{1-x}$F$_{x}$ film at 9\,T (see Fig.\,\ref{rt}). Additionally, a kink appears in $\rho(T)$) at temperatures around 18\,K. The inset in Fig.\,\ref{la} shows $dV/dI$ of LaFeAsO$_{1-x}$F$_{x}$ -- Cu PC with Andreev-reflection minima for varying temperature along with gap behavior evaluated by fitting of the $dV/dI$ curves. Similar as for the SmFeAsO$_{1-x}$F$_{x}$ film
only small gap minima are found.
Again, the critical temperature obtained by extrapolating the BCS curve to $\Delta$=0 is low, i. e. about 18\,K (close to the kink in $\rho(T)$), while superconducting features in $dV/dI$ are seen even at 25\,K as shown in Fig.\,\ref{la}. We attribute this behavior with
the variation of the superconducting properties in the PC region due to the short coherence length and the sample
inhomogeneity.
\begin{figure} [t]
\vspace{3cm}
\begin{center}
\includegraphics[width=8cm,angle=0]{sfa6.eps}
\vspace{-3cm}
\end{center}
\vspace{0cm} \caption[] {(Color on-line) Temperature dependence of the superconducting gap $\Delta$ (triangles)
from the fitting of $dV/dI$ curves shown in the inset. The position of minima in $dV/dI$ is marked by stars.
The solid line represents the BCS-like gap behavior. Inset: $dV/dI$ curves of a
LaFeAsO$_{1-x}$F$_{x}$ -- Cu contact ($R$ = 10\,$\Omega$)
for varying temperature. } \label{la}
\end{figure}
\section{Conclusion}
We investigated the superconducting energy gap in $Re$FeAsO$_{1-x}$F$_{x}$ ($Re$=La, Sm) films by PCAR spectroscopy.
The mean value of the reduced superconducting gap 2$\Delta/kT_c^*$ is found to be close to the BCS value of 3.5
what is in agreement with PCAR data reported in \cite{Chen}. A second larger gap feature is hardly to resolve on PCAR spectra.
The reason can be strong interband quasiparticle scattering, which smears out the larger gap structure in the electronic
DoS both in the tunneling \cite{Noat,Fasano} and in the PCAR spectra. Note also that all superconducting features in $dV/dI$,
including these of non-Andreev nature, have disappeared above $T_c^{onset}\simeq 34\,K$.
We demonstrate that the asymmetry of $dV/dI(V)$ curves is independent of the shape of the curves or whether superconducting
or AR features are present or not in $dV/dI(V)$. Therefore, the asymmetry is not related to the spectroscopy but
reflects bulk properties of $Re$FeAsO$_{1-x}$F$_{x}$. Very probably the asymmetry is caused by the thermoelectric effect.
\section*{Acknowledgements}
Two of the authors (Yu.G.N. $\&$ O.E.K.) would like to thank the Alexander von Humboldt Foundation for the support
and Dr. K. Nenkov for the technical assistance. S.H. and M.K. would like to acknowledge the financial support by DFG
(under project HA5934/1-1).
|
\section{Introduction}
Recently \cite{basko2006metal,basko2006problem} it has been pointed out that the phenomenon of Anderson localization \cite{anderson1958absence}, usually associated with single-particle hopping in a random potential, can be present even in the many-body eigenstates of an interacting quantum system and manifest itself as a phase transition at finite and even infinite temperature.
This phenomenon has been dubbed \emph{many-body localization} (henceforth MBL) and it can be conceived as an example of AL on configuration space, rather than real space. As the geometry of configuration space for a many-body system is quite different from that of a regular lattice in few dimensions, MBL is thought to have properties distinct from those of the single-particle AL.
MBL should be responsible, among other things, of the exact vanishing of the DC conductivity of metals below a critical temperature \cite{basko2006metal} and of the failure \cite{altshuler2010anderson, knysh2010relevance} of the simplest version (and possibly of all versions) of the quantum adiabatic algorithm \cite{farhi2001quantum} for the solutions of NP-complete problems; it has also been studied in disordered Heisenberg spin-chains \cite{brown2008quantum, pal2010many-body} where the phase transition has been linked to the infinite-randomness fixed point. The similarity of some features of MBL to the glass transition in spin and configurational glasses makes it the closest to a \emph{quantum} analog of a glass transition, where the assumptions of equilibrium statistical mechanics fail.
As we said, in some problems MBL is found in typical many-body states\cite{oganesyan2007localization}, namely states sampled with uniform distribution from the spectrum (therefore corresponding to infinite temperature). These states are difficult to study directly, much more than the ground states for which many approximations (DMRG, MPS etc.) can be devised: indeed the only strategy here seems to be exact diagonalization (as used in \cite{pal2010many-body} for example), the exponential complexity of which limits the size of the systems to less than 20 spins; alternatively the study of correlation functions with time-dependent DMRG was used, whose failure to converge due to growing entanglement can signal the onset of delocalization\cite{znidaric2008many}. Analytic results have mainly been obtained by studying the behavior of the perturbation theory for increasing system sizes \cite{knysh2010relevance}.
In this work we report the results of our numerical study on the structure of typical states of a fully-connected quantum spin model with quenched disorder (the Richardson model \cite{richardson1963restricted, richardson1964exact, dukelsky2004colloquium}) which has been introduced as a model of nuclear matter and has been studied in connection with the finite-size scaling of the BCS theory of superconductivity. Integrability allows us to go to sensibly higher spin numbers ($N=50$ spins for single states and we will collect extensive statistics up to $N=40$) and therefore make some educated guesses on the thermodynamic limit of the system.
We will begin by discussing the method we devised for the solution of the Bethe-ansatz equations (the Richardson equations) which is at variance with respect to those very refined ones, used in the study of ground state and low-lying excited states \cite{faribault2009bethe, faribault2008exact} (our method will be close in spirit to that used in \cite{faribault2011numeric}, which appeared while this work was being completed).
Then we will discuss the observables, since we will face the problem that the classical observables in localization studies, the inverse participation ratio (IPR), is computationally heavy (its complexity goes as $2^N$, although still smaller than $\Ord{2^{3N}}$ steps required by exact diagonalization).
We devised a Montecarlo method for the measure of IPR and performed an extensive study of an Edwards-Anderson order parameter $q$ \cite{mezard1987spin}, which is related to the average (1,$N$-1)-entanglement (the Meyer-Wallach\cite{meyer2002global} entanglement measure) and to the average Hamming distance $L$ between states in the computational basis whose superposition forms the eigenstate. As we show, $q$ is related to the long-time spin relaxation and therefore thermalization is not achieved as long as $q>0$.
We also observe that for the Richardson model, $q$ is in 1-to-1 correspondence with the IPR ${\mathcal I}$ (but the relation is different from what found for disordered Heisenberg model\cite{viola2007generalized}). We find for the thermodynamic limit of $q$ the deceptively simple expression as a function of the hopping $g$: $q=(1+g)^{-1}$ which implies in the same limit, for the average Hamming distance $L/N=\frac{g}{2(1+g)}$. We will define a local entropy density $s=\log{\mathcal I}/2L$, for which we find numerically a well-defined thermodynamic limit, although the limiting form does not seem to have the simple character of the previous quantities.
We have also studied the clustering properties of the eigenstates, and we have not found any presence of clusters but rather as the hopping $g$ is increased the eigenstates spread rather uniformly over the configuration space. Neighboring states in energy have very close values of $q$ but their overlap (a measure common to spin-glass studies) is close to 0. This means that clusters can be formed if one takes a superposition of states in a small energy interval to make a microcanonic density matrix.
The picture that emerges from this analysis is that there is no many-body localization-delocalization phase transition in this model{\color{blue}:} although the states appear de-localized on the computational basis for any finite $g$, the average single-spin observables are always localized.
Finally, we discuss the role of integrability in the previous predictions and the implications of our findings for more natural cases where integrability is broken\cite{canovi2010quantum}.
\section{The model and its solution}
The Richardson model \cite{richardson1963restricted, richardson1964exact} is an XX-model (i.e.\ with no $s^zs^z$ coupling) of pairwise interacting spins with arbitrary longitudinal fields
\begin{equation}\label{hamiltonian}
H=-\frac{g}{N}\sum_{\alpha,\beta=1}^N s^+_\alpha s^-_\beta-\sum_{\alpha=1}^N h_\alpha s^z_\alpha,
\end{equation}
where $s^{x,y,z}$ are spin-$\frac{1}{2}$ representation of $SU(2)$ algebra. This model can accommodate quenched randomness in the arbitrary choice of the fields $h_\alpha$.\footnote{It is known that this model in 1-d reduces to non-interacting fermions and hence localizes for arbitrary small disorder. This is a peculiarity of 1-d and is due to the existence of the Jordan-Wigner map.} We choose a Gaussian distribution for them, with $\mean{h}=0, \mean{h^2}=1$. First of all one notices that the total spin $S^z$ is conserved and we focus on the subspace $S_z=0$ which exists only for even $N$. All the states in the sector $S^z=(2M-N)/2$ can be found by applications of $M$ generalized raising operators on the state with all spin down:
\begin{equation}\label{eigenstates}
\ket{E[w]}= \prod_{j=1}^M B(w_j) |\downarrow\ldots\downarrow\rangle,
\end{equation}
where the $M$ \emph{Richardson roots} $w_j$ satisfy the $M$ \emph{Richardson equations}:
\begin{equation}\label{RE}
\forall j=1,...,M:\quad
\frac{N}{g}+\sum_{\alpha=1}^N\frac{1}{w_j-h_\alpha}-\sum_{k=1, k\neq j}^M\frac{2}{w_j-w_k}=0
\end{equation}
in terms of which the raising operators are
\begin{equation}\label{B}
B(w)=\sum_{\alpha=1}^N\frac{S_\alpha^+}{w-h_\alpha}
\end{equation}
and the energy of the state is given by
\begin{equation}\label{nrg}
E[w]=\sum_{j=1}^M w_j-\sum_{\alpha=1}^N \frac{h_\alpha}{2}.
\end{equation}
As we said, we will focus on $S_z=0$ so we will have $M=N/2$, which means that we have to solve $N/2$ coupled nonlinear equations, the different $\binom{N}{N/2}$ states are determined by the boundary conditions as $g\to 0$. Indeed as $g\to 0$ the roots tend to some of the fields $h_\alpha$ and the choice of the set can be used to label the state at any $g$.
Starting from $g=0$ instead, and increasing $g$, the roots start departing from their initial $h$'s values, collide and become complex conjugate. By increasing $g$ sometime they recombine and return real. The number of roots that eventually ($g\to\infty$) diverge is equal to the total spin $S$ of the state (which is a conserved quantum number at infinite $g$). An algorithm which can follow the evolution of the roots with $g$ has to take into account these changes in the nature of the solution, where the roots become complex conjugate. These critical points, for random choices of the $h$'s can occur at particularly close values of $g$ and this can create troubles for the algorithm.\footnote{This problem is not so serious for the ground state and first excited states so one can go to much higher values of $N$ without losing accuracy.} In order to pass the critical points a change of variable is needed, and one natural choice is \cite{faribault2009bethe}:
\begin{eqnarray*}
&& w_+ = 2 h_c-w_1-w_2 \\
&& w_- = (w_1-w_2)^2
\end{eqnarray*}
in which $w_1$ and $w_2$ are the root colliding on the level $h_c$.
When more than a pair of roots collide in a too small interval of $g$ this change of variables may not be sufficiently accurate and one should think of something else (if one does not want to reduce the step in the increment of $g$ indefinitely). The most general change of variables which smooths out the evolution across critical points is that which goes from the roots $w_j$ to the coefficients $c_i$ of the characteristic polynomial $p(w)$ --i.e.\ the polynomial whose all and only roots are the $w_j$'s.
This polynomial is quite interesting in itself as it satisfies a second order differential equation whose polynomial solutions have been classified by Heines and Stjielties \cite{szego1939orthogonal}.\footnote{The equation is $-h(x)p''(x)+\left(\frac{h(x)}{g}+h'(x)\right)p'(x)-V(x)p(x)=0,$ where $h(x)=\prod_{\alpha=1}^N(x-h_\alpha)$, $V(x)=\sum_{\alpha=1}^N\frac{h(x)A_\alpha}{x-h_\alpha}$. The problem to be solved is to find a set of $A_\alpha$'s such that there exists a polynomial solution of this equation. The solution will automatically satisfy also $A_\alpha=\frac{p'(h_\alpha)}{p(h_\alpha)}.$ A similar approach has also been investigated in the recent work \cite{faribault2011numeric}.} Following the evolution of the coefficients $c_i(g)$ is a viable alternative to following the roots but we found out that the best strategy is a combination of both evolutions. Therefore we follow the evolution of the roots, extrapolating the coefficients and using them to correct the position of the roots at the next step in the evolution. In this way we do not implement any change of variables explicitly and we do not have to track the position of critical points. This algorithm
\footnote{Python code is available on the webpage: \\
http://www.sissa.it/statistical/PapersCode/Richardson/} can be used on a desktop computer to find the roots of typical states with about 50 spins, although in order to collect extensive statistics we have limited ourselves to $N=40$.
\section{Order parameters}
\subsection{IPR, entanglement, average Hamming radius of an eigenstate and a local entropy}
Once obtained the roots one is faced with the task of studying the state. The quantity characterizing the localization/delocalization properties of a state (on the basis of $s^z_i$'s: the computational basis or configuration space ${\mathcal C}$) is the inverse participation ratio of an eigenstate $\ket{E}$:
\begin{equation}
{\mathcal I}=\left(\sum_{{s_1,...,s_N,\ \sum_\alpha s_\alpha=0}}|\braket{s_1,...,s_N}{E}|^4\right)^{-1}.
\end{equation}
We will see that for all $g>0$, $\log{\mathcal I}\propto N$, i.e., an exponential number of sites of the hypercube of spin configurations is covered;
nevertheless, we generally observe that
\begin{equation}
\lim_{N\to\infty} \frac{{\mathcal I}}{\binom{N}{N/2}}\to 0,
\end{equation}
which flags instead a single-particle \emph{localized} phase, according to the definition common in AL studies. In fact, the analysis of single-particle observables will confirm this scenario.
The amplitudes $\braket{s_1,...,s_N}{E}$ can be calculated as ratio of determinants of $(N/2)\times (N/2)$ matrices (therefore in time $\sim N^3$) once the roots $w_j$ are known. However the number of terms in the sum is exponential in $N$ so the calculation of ${\mathcal I}$ requires an exponential number of terms\footnote{We have looked for a shortcut to evaluate this sum but to our knowledge integrability does not help us here.} and we are limited again to twenty spins or so.
We found two ways around this difficulty: they are complementary and can be checked one against the other for consistency. First, we devised a Montecarlo algorithm for the evaluation of ${\mathcal I}$. Define the probabilities $p_a=|\braket{a}{E}|^2$ where $a\in{\mathcal C}$ stands for one of the $\binom{N}{N/2}$ allowed configurations of spins which constitute the configuration space ${\mathcal C}$. We perform a random walk with the probabilities $p_a$'s, namely start from a random configuration $a$ and we try to move to a random one of the $(N/2)^2$ neighboring states, say $b$, by accepting the move with probability $\min(1,p_b/p_a)$. This involves only one computation of $p_b$, which takes time $\sim N^3$. The random walk proceeds in this way, generating a history of configurations $a$ for which we can take the average over Montecarlo time of $p_a$. The inverse of this value gives ${\mathcal I}$.
The intensive quantity is $\log{\mathcal I} /N$, which can then be averaged over different states and realizations. We observe that for all $g=\Ord{1}$ the value of $\ln{\mathcal I}\propto N$, testifying then that each state occupies an exponential number of states in the configuration space.
The second method is to find another quantity which can be computed in polynomial time and to link it to ${\mathcal I}$. Since the average values $\bra{E}s_\alpha^z\ket{E}$ can be expressed again in terms of determinants they can be calculated in $\Ord{N^3}$ time: therefore one is led to consider a \emph{microcanonical} version the Edwards-Anderson order parameter associated to a single eigenstate
\begin{equation}
q(E)=\frac{4}{N}\sum_{\alpha=1}^N\bra{E}s_\alpha^z\ket{E}^2,
\end{equation}
with this normalization $q\in[0,1]$. The average over eigenstates is
\begin{equation}
q=\frac{1}{2^N}\sum_E q(E).
\label{eq:avgqEA}
\end{equation}
To get the physical significance of this quantity, following \cite{pal2010many-body} we start with a slightly magnetized spin $\alpha$ in an infinite temperature state:
\begin{equation}
\rho_0=({\mathbb I}+\epsilon s^z_\alpha)/2^N
\end{equation}
with magnetization $\langle s^z_\alpha\rangle_0=\Tr{\rho_0 s^z_\alpha}=\epsilon/4$ (as $s_z^2=1/4$). The same magnetization at large time $t$ in the diagonal approximation reads
\begin{equation}
\langle s^z_\alpha\rangle_\infty=\lim_{t\to\infty}\Tr{e^{-iHt}\rho_0e^{iHt}s^z_\alpha}= \frac{\epsilon}{2^N}\sum_E\bra{E}s^z_\alpha\ket{E}^2.
\end{equation}
Therefore, averaging over $\alpha$ we obtain the equality with eq.\ (\ref{eq:avgqEA}):
\begin{equation}
q=\frac{1}{N}\sum_{\alpha}\frac{\langle s^z_\alpha\rangle_\infty}{\langle s^z_\alpha\rangle_0},
\end{equation}
namely the previously defined EA order parameter is the average survival fraction of the initial magnetization after very long times.
We notice two more things:\cite{viola2007generalized} one, that $q$ is related to the average purity of the state (here we use the total $S^z=0$):
\begin{equation}
q=\frac{2}{N}\sum_{\alpha}\Tr{\rho_\alpha^2}-1
\end{equation}
and two, that $q$ is related to the average Hamming distance of the points in configuration space when sampled with the probability distribution $p_a$:
\begin{equation}
d_{a,b}=\sum_{\alpha=1}^N\frac{1-4\bra{a}s_\alpha^z\ket{a}\bra{b}s_\alpha^z\ket{b}}{2},
\end{equation}
and multiplying by $p_{a},\ p_{b}$ and summing over $a,b$ we find:
\begin{equation}
\label{eq:qd}
L \equiv \avg{d}=\frac{N}{2}(1-q).
\end{equation}
So $q$ is computationally easy and it captures both some geometric properties of the covering of the configuration space by an eigenstate and the long-time correlation function for $s^z$. We averaged $q$ over the spectrum (sample over typical states) and then over realizations (the number of which depends on the size of the system but it will never be less than 100).
We found this average $\mean{q}$ as a function of $g$ for $g\in[0,40]$ and $N=16,...,38$ and studied the pointwise finite-size scaling (in the form $q_N(g)=q(g)+c_1(g)/N+c_3(g)/N^3$) to obtain the thermodynamic limit of $q$ (see Figure \ref{fig:qg}). We fit the data using a ratio of polynomials with the condition that $q(0)=1$ and we found that averaging over the state and the realization of disorder
\begin{equation}
\label{eq:qgpade}
\mean{q}=\frac{1+3\times 10^{-8}g}{1+1.003g+0.009g^2}\simeq\frac{1}{1+g}
\end{equation}
works in the whole range of data to an error of at most $0.5\%$. We therefore conjecture this to be the correct functional form of the EA order parameter at infinite temperature.
\begin{figure}[htbp]
\centering
\includegraphics[width=8cm]{./qglimit.pdf}
\caption{The pointwise extrapolation of the function $q$ as a function of $g$. The fit $q=1/(1+g)$ is not distinguishable from the data.}
\label{fig:qg}
\end{figure}
We can now go back to the relationship between the IPR and $q$, better expressed as a relation between $\log{\mathcal I}$ and $L$. We notice a one-to-one correspondence between average values these two quantities already at the level of second-order perturbation theory in $g$ starting from a given state with $N/2$ spins up ${\mathcal S}_\uparrow$ and $N/2$ spins down ${\mathcal S}_\downarrow$:
\begin{equation}
{\mathcal I} = 1 + \frac{2 g^2}{N^2} A + o(g^2)
\end{equation}
where we defined a sum over pairs of up and down spins of the given state:
\begin{equation}
A = \sum_{\alpha \in {\mathcal S}_{\uparrow},\, \beta \in {\mathcal S}_{\downarrow}} \frac{1}{ (h_\alpha - h_\beta)^2}
\end{equation}
Since $A$ is dominated by small denominators, it will be typically $A = O(N^4)$ and therefore from the expression for IPR we see the perturbative regime is valid for $g \ll1/N$. With an analogous computation we get:
\begin{equation}
\label{Qperturbation}
L = \frac{4 g^2}{N^2} A + o(g^2)
\end{equation}
Eliminating $g$ between the two relations and using (\ref{eq:qd}) one obtains, independently of the state and of the quenched randomness (therefore the relation holds also on average):
\begin{equation}
\label{ipdrel}
\log {\mathcal I} \simeq \frac{L}{2}.
\end{equation}
So the relation is linear for small $g$. To see how this relation is modified at higher values of $g$ we have again to resort to numerics. From the data it is clear that a strict relation exists between $\mean{\log{\mathcal I}}$ and $\mean{L}$ as one can see in Figure \ref{fig:IPRL}
\begin{figure}[htbp]
\centering
\includegraphics[width=8cm]{LogIPRL.pdf}
\caption{$\log {\mathcal I}$ as a function of the average distance $L$. The points are (blue, pink, yellow) $N = 28,30,32$ averaged over 100 realizations: the dashed straight line is the second order perturbation theory approximation Eq.\ (\ref{ipdrel}).}
\label{fig:IPRL}
\end{figure}
By using the previous Montecarlo calculation for ${\mathcal I}$ we can plot $\log{\mathcal I}$ vs.\ $L$, showing that the relation is almost linear. The degree of non-linearity is measured by the ratio
\begin{equation}
s=\frac{\log{\mathcal I}}{2L}
\end{equation}
which can be interpreted as a local entropy\cite{campbell1987random}. In fact, $2L=N(1-q)$ can be interpreted as the number of free spins (whose value of $s^z$ is close to 0) while ${\mathcal I}$ is the number of configurations. If we want $2L$ spins to be responsible to ${\mathcal I}$ states then each of these spins should account for a degeneracy of $e^s$, from which the interpretation as an entropy density.
The distribution of $L$ over states and realizations becomes more and more peaked as $N$ grows, since we observe the variance $\delta L^2\propto N$. The same occurrs to $\log{\mathcal I}$, whose variance goes $\propto N$ in the region of $g$ considered. Therefore the average value of $s$ becomes typical in the large-$N$ limit.
\begin{figure}[htbp]
\centering
\includegraphics[width=8cm]{sg-fit.pdf}
\caption{Local entropy as a function of $g$. The points are $N=28,32,36$ (blue, pink, yellow) all averaged over 100 realizations. The fit is a (1,1)-Pade' approximation conditioned to $s(0)=1/4$.}
\label{fig:sg}
\end{figure}
In the curves of Figure \ref{fig:sg} the entropy $s$ grows from the value of $1/4=0.250$ predicted by perturbation theory to an asymptotic value of $s=0.383\pm0.003$.\footnote{In the figure we show also a rational function best fit $s(g)=0.403\ \frac{0.452+g}{0.656+g},
$ which has however an error of $5\%$ in the asymptotic value $s(\infty)=0.403$ instead of $0.383$, the value obtained by averaging on many more realizations and including smaller $N$ in the fit.} This value is not what one would expect from a uniform superposition over $\binom{N}{N/2}$ states, since in that case $L=N/2$, $\log{\mathcal I}\simeq N\log 2$ and the familiar value $s=\log 2=0.693$ is roughly \emph{twice} as much as we expect. This leads us to think that the most probable structure of the delocalized state at increasing $g$ still retains a pair structure. We can build a toy model of delocalization in the typical eigenstates, by assuming that $Nq$ spins are localized on their $g=0$ values and that the remaining $N(1-q)$ spins are instead divided into couples, where couples are formed between almost resonating spins of opposite orientation. The couples are in one of the random valence bond states $\ket{\uparrow\downarrow}\pm\ket{\downarrow\uparrow}$ which are indeed the two $S_z=0$ eigenstates of the two-body Hamiltonian
\begin{equation}
H_2=-\frac{g}{N}(s^+_1s^-_2+s^+_2s^-_1)-h (s^z_1+s^z_2)
\end{equation}
where we have assumed that $h_1\simeq h_2=h$.
This predicts that
\begin{equation}
{\mathcal I}\sim2^{N(1-q)/2}=2^L
\end{equation}
and so that we should have a constant entropy $s=\log(2)/2=0.347$, slightly smaller than the observed value at large $g$ and off by 40\% at small $g$. The pair structure of a given eigenstate can be observed in Figure \ref{fig:magspeed} where we plot the values of $m_n^2\equiv\bra{E}s_n^z\ket{E}^2$ for a given eigenstate $\ket{E}$, ordering the spins by increasing $h_n$ (so that almost-resonant spins are nearest neighbors). We see a clear valence bond-like pair correlation in the values of the squared magnetization.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{magspeed}
\caption{Squared magnetizations for increasing $g$ (from white to black), where the spins are ordered by increasing magnetic field $h_n$. A consistent number of valence bond pairing is observed as a good fraction of neighboring spins (e.g. $n=8,9$, $n=15,16$ and $n=31,32$) have the same speed.}
\label{fig:magspeed}
\end{center}
\end{figure}
With the available data, we can discuss issues like the presence of multiple clusters in the same energy level $\ket{E}$. In fact, by randomly restarting the Montecarlo routine with the same $p_a$'s if multiple clusters exist, we would expect to sample them according to their basin of attraction. Moreover we can rely on analytic results (such as those for $q$) to compare the Montecarlo averages with: clusterization and ergodicity-breaking would translate in a difference between these two results (as the random-walk would get stuck in a cluster and would not explore the whole configuration space). In the region where we can trust our numerics ($ g \gtrsim 0.02 $) Montecarlo averages converge to the analytic results, though a slowdown of the dynamics is observed (see below).
\subsection{Dynamics of Montecarlo and other quantities}
The Montecarlo routine which allows for importance sampling of the distribution $p_a$ allows other measures of the geometry of the state. We can now study the similarities between the dynamics of importance sampling on $p_a$ and that of random percolation on the hypercube, which has been proposed as a model of relaxation in a glassy system \cite{campbell1987random}. We will find that in both cases, a stretched exponential is the best fit and that the exponent depends on the coupling constant $g$. This, we believe, is a remarkable similarity.
An important quantity in this sense is the time dependence of the average distance from the starting point. Consider the Hamming distance $H(t)$ ($t$ is Montecarlo time) from the starting point $H(t)=|a(t)-a(0)|$. For $t \gg 1$, $H(t)$ is fit quite accurately by a stretched exponential ansatz of the form:
\begin{equation}
H(t) = L \left(1 - e^{-\left(\frac{t}{\tau}\right)^\beta} \right),
\end{equation}
where $L$ is the average distance introduced before and $\beta$ is a new characteristic exponent.
Let us consider the behavior of the exponent $\beta$ with respect to $g$, as plotted in Fig.\ \ref{fig:betag}. Even if the results become quite noisy for small $g$, we can still see that for small values of $g$, $\beta$ stays close to $1$, while as $g$ increases, $\beta$ decreases, although quite slowly. Instead for the time-scale $\tau$ we find, apart from the monotonic decrease with $g$, which is to be expected on general grounds that, for $g\gtrsim 1$, $\tau\propto N^{3/2}$ which we propose without explanation.
\begin{figure}[htbp]
\centering
\includegraphics[width=8cm]{betag.pdf}\\
\includegraphics[width=8cm]{taug.pdf}
\caption{\emph{Up.} The stretched exponential exponent $\beta$ data as a function of $g$ for $N=28,32,36$ (blue, pink, yellow) together with a fit of the form $(1+a_1 g)/(b_0+b_1 g+b_2 g^2)$. \emph{Down.} The timescale $\tau$ as a function of $g$ for $N=18,20,...,34$ (red to green). The scaling $\tau\propto N^{3/2}$ is evidently good, in particular in the region $g>1$.}
\label{fig:betag}
\end{figure}
The small time behavior of $H(t)$ can be used to obtain some information about the local structure of the state. In particular we can set
\begin{equation}
k \equiv \frac{H(1)}{2} = \frac{4}{N^2} \sum_{\langle a, b \rangle} \min (p_a, p_b)
\end{equation}
where the last equality follows from the Montecarlo rate and the sum is over nearest-neighbour states. This quantity can be considered as a measure of the local connectivity, that is, the average fraction of active links.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{kg.pdf}
\caption{Local connectivity as a function of $g$. Different lines corresponds to $N = 18,24,30$ (blue, pink, yellow)}
\label{fig:conng}
\end{center}
\end{figure}
From Fig.\ \ref{fig:conng}, we may deduce two things: one is that the connectivity stays well below $1$ even for large $g$, confirming, as we claimed before, that the typical state is never uniformly spread over the hypercube; the second is that the connectivity scales with $N$ as $N^{-1}$ for small $g$ and with $N^{-1/2}$ for large $g$ (a fit $k=A/N^\alpha$ shows a continuously decreasing $\alpha$ from 1 to $1/2$). Since the number of diverging roots is proportional to the total spin of the eigenstate at infinite $g$, and since the more roots diverge the more equally distributed the terms of each creation operator $B$ are, we can argue the state will be better spread for larger total spin $S$. As for large $N$, the total spin of a typical state will increase only as $\sqrt{N}$ this explains the depletion of the local connectivity.
\subsection{Some perturbation theory}
Let us see what the predictions of perturbation theory are for our system. Let us start at $g=0$ from state $a$, with energy $E_{a}$. The states at distance 2 from $a$ have energies
\begin{equation}
\Delta^{(2)}=E_{b}-E_{a}=h_\alpha-h_\beta
\end{equation}
where the couple $(\alpha,\beta)\in {\mathcal S}_\uparrow\times{\mathcal S}_\downarrow$ defines the spins which have been flipped up and down in going from $a$ to $b$. The typical value of $\Delta^{(2)}$ is $\sqrt{\avg{(\Delta^{(2)})^2}}=\sqrt{2}=\Ord{1}$ however the minimum value is $\Ord{N^{-2}}$ which we write $x^{(2)}/N^2$ where $x^{(2)}=\Ord{1}$. So the corresponding term in perturbation theory for the wave function is
\begin{equation}
A_{b}=\frac{(g/N)}{x^{(2)}/N^2}=\frac{gN}{x^{(2)}}.
\end{equation}
In this way we can go on at arbitrary distance $2k$, to the state $b^k$, the amplitude thus having $k$ denominators of $\Ord{1/N^2}$
\begin{equation}
A_{b^{k}}=\frac{(gN)^k}{x^{(2)}x^{(4)}...\ x^{(2k)}},
\end{equation}
where $x$ are random variables of $\Ord{1}$. For any given $a$ there are only $\Ord{1}$ neighboring states with $\Delta\sim 1/N^2$, so the number of such $b^k$ states at distance $2k$ from $a$ is $\Ord{1}$ out of $N^{2k}$ (also the number of relevant paths does not grow as $k!$). These can be called a \emph{direct} or \emph{percolating} contribution. However already at distance 4 we observe another type of contribution, which one is tempted to dub a \emph{tunnelling} contribution, in which although the final denominator $\Delta^{(4)}=h_\alpha-h_\beta+h_\gamma-h_\delta=z^{(4)}/N^4$ each of the two paths leading to the minimum $h_\alpha-h_\beta\simeq -(h_\gamma-h_\delta)=y^{(2)}=\Ord{1}$, where $\alpha,\gamma\in{\mathcal S}_\uparrow$ and $\beta,\delta\in{\mathcal S}_\downarrow$. Again this contribution is of order:
\begin{equation}
A_{b}=\frac{(g/N)}{y^{(2)}}\frac{(g/N)}{z^{(4)}/N^4}=\frac{(gN)^2}{y^{(2)}z^{(4)}},
\end{equation}
while the amplitudes corresponding to the distance-2 intermediate steps are $\Ord{1/N}$.
The distribution of $x,z$ can be found by using the theory of extreme value statistics \cite{kotz2000extreme}, while $y$'s are typical values of field differences and none of these distribution depends on $N$. We will stop here our analysis of perturbation theory as this would require a separate work by itself. It is sufficient for us to notice that only the combination $gN$ appears in all terms of the series so scaling $g$ to zero like $1/N^{1+\epsilon}$ for every $\epsilon>0$ each term would go to 0 and the series would trivially converge unless the series is asymptotic in $gN$. Notice that an argument based on a Bethe-lattice approximation for the configuration space ${\mathcal C}$ (see \cite{abou1973selfconsistent}) would give $g_c\sim 1/N \log N$ for the localization transition.
\subsection{Setup of an exponential IPR}
From perturbation theory (and from the Bethe-lattice approximation result \cite{abou1973selfconsistent}) one could argue that, if a phase transition occurs, it is at $g\sim 1/N$ (an extra factor $1/\log N$ would not be noticed for our moderately large $N$). But does a phase transition in the geometric properties of the eigenstate occur?
First we analyze the quantity for which we have more extensive statistics (because of his polynomial complexity), $L$. A phase transition in $L$ would mean that, set $\gamma=g N$, there exists a $\gamma_c>0$ such that for $\gamma<\gamma_c$, $L/N\to 0$ and for $\gamma>\gamma_c$, $L/N\sim(\gamma-\gamma_c)^\delta$ where $\delta$ is a critical exponent. We have analysed our data for $g>0.01$ and $g<0.2$ and we can conclude that this is not the behavior observed. The behavior is more consistent with $\gamma_c=0,\ \delta=1$ or with a crossover, in which the limit $L/N$ when $N\to\infty$ is a smooth function of $g$ which vanishes at $g=0$. The matching with the part of the curve at finite $g$ is smooth and the limiting behavior is as described before.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{./LogIPRScaling.pdf}
\caption{The logarithm of the inverse participation ratio divided by the size for $N$ between $8$ and $18$, computed exactly and averaged over states and realizations. This graph shows no hint of a phase transition at $g\lesssim 1/N$.}
\label{fig:xactlogipr}
\end{center}
\end{figure}
There is the possibility however that although $L\sim N$ always, we have two phases: $\log{\mathcal I}\sim 1$ and $\log{\mathcal I}\sim N$ between which a transition occurs. This could happen if an eigenstate spread along one (or a few) directions without covering an exponential number of spin configurations. We have excluded this by both direct analysis of $\log{\mathcal I}/N$ data and by the observation that the relation between $\log{\mathcal I}$ and $L$ remains valid all the way to small $g$ (small here means $g\lesssim 1/N$). As $\log {\mathcal I}/N$ becomes soon independent of $N$ without any scaling of $g$ needed (see Fig.\ \ref{fig:xactlogipr}) we are led to conclude that no phase transition occurs as the system occupies an exponential number of sites of the computational basis for any $g>0$.
\section{Breaking of integrability}
By considering the Richardson model essentially as a hopping process on the hypercube with random site energies given by the unperturbed energies we have found that the eigenstates are always covering an exponential number of spin configurations but nonetheless $q>0$ meaning thermalization is not achieved.
From this point of view it is not obvious which role, if any, integrability plays.
But, on the other hand, one would infer that the integrability of the model must play an essential role beyond providing the methods used for its solution. In particular, if the integrals of motion are too much local, integrability can have the effect of freezing the expectation values of local quantities. To support such a claim we investigated a very similar non-integrable model, in which the hopping coefficients are not uniformly equal to $g$ as in (\ref{hamiltonian}) but instead $N(N-1)/2$ random variables $g_{\alpha,\beta}=g(1+\epsilon\eta_{\alpha,\beta})$ where $\eta_{\alpha,\beta}=\eta_{\beta,\alpha}=\pm 1$ with probability $1/2$. Randomness in the fields is retained. The hamiltonian is
\begin{equation}
H=-\frac{g}{N}\sum_{\alpha,\beta}(1+\epsilon\eta_{\alpha,\beta})s^+_\alpha s^-_\beta-\sum_{\alpha}h_\alpha s^z_\alpha.
\label{eq:Hnoni}
\end{equation}
We observe a decrease of the value of $q$ (averaged both over $E,\eta$ and $h$) as expected, in particular for sufficiently large $g$ we are confident to say that $q\to 0$ for $N\to\infty$ and the system becomes ergodic.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=7.5cm]{./scaling0104.pdf}
\includegraphics[width=7.5cm]{./fit4104.pdf}
\caption{Averaged microcanonical $q$ for the non-integrable Hamiltonian eq.\ (\ref{eq:Hnoni}) for \emph{up:} $g=0.1,\ \epsilon=0.4$ and \emph{down}: $g=4.1,\epsilon=0.4$. The data in the lower panel are fit by a power law $aN^{-\gamma}$ with $a=290$ and $\gamma\simeq 4$. The exponent $\gamma$ seems to be $g$-dependent.}
\label{fig:qnonin}
\end{center}
\end{figure}
For small $g$ however the situation is not so clear. The limit $N\to\infty$ could actually be zero or not, what is clear is that the $N$-dependence is not settled (compare the upper and lower panel of Figure \ref{fig:qnonin}) for $N=16$, the largest system size that we can attain. This leads to two competing scenarios: in the first we have ergodicity as soon as $\epsilon>0$; in the second, one could identify a finite $g_c(\epsilon)$ such that for $g<g_c$, $q>0$ and for $g>g_c$ we have $q=0$ in the thermodynamic limit. The latter would have a MBL transition at the said $g_c$. Much more extensive numerical work is needed to decide between these two scenarios. We leave the resolution of this issue for the future.
\section{Conclusions and some directions for further work}
We have performed a numerical study of typical states of the Richardson model with quenched disorder (an example of Gaudin magnet and an integrable system). We have found no evidence of a delocalization phase transition although typical eigenstates occupy an exponential number of states in the basis of $s^z_i$'s for any $g>0$.
We have devised a method to calculate the IPR without summing over exponentially many states and studying its connections with a microcanonical version of the Edwards-Anderson order parameter, which measures the fraction of surviving magnetization at infinite temperature and for long times. Of this order parameter, we have conjectured the thermodynamic limit at infinite temperature as $q=1/(1+g)$. We were unable to obtain the temperature dependence of this quantity, as sampling from the Boltzmann distribution is not straightforward within our framework.
For what concerns the absence of a MBL \emph{phase transition} we can point out two peculiarities of our system as responsible for its absence. One is integrability and the other is the infinite range of the Hamiltonian. We have therefore studied small-size systems (up to $N=16$ spins) with an extra integrability breaking term of size $\epsilon$. We observe a sharp reduction of $q$, which in some range of parameters could lead to think to a phase transition where $q=0$ for $g>g_c(\epsilon)$. However it is possible that in the complementary region $(g<g_c(\epsilon))$ the decrease with $N$ starts from a value of $N$ impossible to reach with our limited numerics so we are unable to see that $q=0$ for all $g$ as soon as $\epsilon>0$.
Unfortunately this dichotomy is unlikely to be settled with the currently accessible values of $N$ and in absence of an established scaling theory of MBL.
We point out that the Richardson model is one of a family of integrable spin systems (generalized Gaudin's magnets, see \cite{ortiz2005exactly}) which can be studied with minor modifications of the methods introduced in this paper. We leave this too for further work.
\section{Acknowledgements}
It is a pleasure for us to thank M.Mueller, G.Santoro, S.Sondhi, A.Silva, X.-Q. Yu and in particular R.Fazio and D.Huse for their suggestions and comments.
|
\section{Introduction}
One of the most remarkable discoveries of the past decade is that
the universe is currently in an accelerating phase, which means that
expanding velocity of the universe is increasing. This phenomenon is
supported by data related to the luminosity measurements of high red
shift supernovae \citep{Per99,Rie98,Ast06,Woo07}, measurement of
degree-scale anisotropies in the cosmic microwave background (CMB)
\citep{Mil99,Han00,Spe03} and large scale structure (LSS)
\citep{Col01,Teg04,Col05,Spr06}. The rigorous treatment of this
phenomenon can be provided in the framework of general relativity.
In the expression of general relativity, late time acceleration can
be explained either by an exotic fluid with large negative pressure
that is dubbed as \textit{dark energy} in literature, or by
modifying the gravity itself which is dubbed as \textit{dark
geometry} or \textit{dark gravity} proposal. The first and simplest
candidate of dark energy is the cosmological constant, $\Lambda$
\citep{Sah00,Pad03,Cop06}. But there are theoretical problems
associated with it, such as its unusual small numerical value ( the
fine tuning problem), no dynamical behavior and its unknown origin
\citep{Wei89,Car01,Cal99}. These problems have forced cosmologists
to introduce alternatives in which dark energy evolves during the
universe evolution. Scalar field models with their specific features
provide an interesting alternative for cosmological constant and can
reduce the fine tuning and coincidence problems. In this respect,
several candidate models have been proposed: "quintessence" scalar
field \citep{Rat88,Sai00,Bra00,Bar00,Sah00,Sah02,Sam03}, phantom
fields \citep{Cal02,Tsu04,Cal05, Cai10, Moy09} and chaplygin gas
\citep{Kam01,Dev03,Ame03,Roo07,Bou07,Zha06a,Zha06b,Ber04,Bie05,Hey08,Roo08a,Roo08b,Zha09,Set09}
are among these candidates.
As an alternative for dark energy, modification of gravity can be
accounted for the late time acceleration. Among the most popular
modified gravity scenarios which may successfully describe the
cosmic speed-up, is $f(R)$ gravity
\citep{Cap03,Sot10,Noj03,Noj04,Noj05,Noj06,Noj07,Noj08,Bam08,Car04,Ame07,Noz08a,Ata08,Saa09,
Set08}. Modified gravity also can be achieved by extra-dimensional
theory in which the observable universe is a 4-dimensional brane
embedded in a five-dimensional bulk. Dvali- Gabadadze-Porrati (DGP)
model is one of the extra-dimensional models that can describe
late-time acceleration of the universe in its self-accelerating
branch due to leakage of gravity to the extra dimension
\citep{Dva00a,Dva00b,Dva01,Dva02,Def01,Lue06}.
Recent observations constrain the equation of state parameter of the
dark energy to be $\omega_{X}\approx-1$ and even $\omega_{X}<-1$
\citep{Mel03,Rie04,Kom09}. One of the candidates for dark energy of
this kind is the phantom scalar field. This component has the
capability to create the mentioned acceleration and its behavior is
extremely fitted to observations. But it suffers from problems; it
violates the null energy condition and its energy density increases
with expansion of the universe which is an unphysical behavior. Also
it causes the quantum vacuum instabilities. So, cosmologists have
tried to realize a kind of phantom-like behavior in the cosmological
models without introduction of phantom fields
\citep{Sah03,Sah04,Sht00,Lue04}. With phantom mimicry ( the
phantom-like behavior), effective equation of state parameter of
dark component remains less than $-1$ and effective energy density
of the unverse increases with cosmic expansion. In the framework of
general relativity these two expressions are equivalent. However, in
the effective picture of cosmological models, in order to satisfy
phantom-like behavior, the mentioned two expressions must be
investigated separately. The phantom mimicry discussed in this study
has a geometric origin. In this paper, we discuss briefly
DGP-inspired theoretical models that realize phantom-like behavior.
We present cosmological dynamics in each proposed model and then we
compare these models with $\Lambda$CDM and also with each other
through investigation of their expansion histories. Within a
dynamical system ( phase space) approach, we show that the
accelerating phase of the universe in the $f(R)$-DGP scenario is
stable if one consider the curvature fluid as a phantom scalar field
in the equivalent scalar-tensor theory, otherwise it is a transient
and unstable phenomenon. As another important probe, we study the
age of the universe in each model. We show that some of these models
account for a transient accelerated phase.
\section{The phantom-like evolution}
As we have pointed out in the introduction, the observational
evidences show that the equation of state parameter of dark energy
in the universe can be less than $-1$. Many attempts have been made
to find dark energy models that allow the so-called phantom dark
energy: a dark energy component with large negative pressure and
negative kinetic energy with equation of state parameter less than
$-1$. A phantom field is described by the following action
\begin{equation}
S=\int \Big[\frac{1}{2}g^{\mu\nu}\partial_{\mu}\varphi\partial_{\nu}
\varphi-V(\varphi)\Big]\sqrt{-g}d^{4}x\,,
\end{equation}
The kinetic energy term of the phantom field in the corresponding
lagrangian enters with opposite sign in contrast to the ordinary
matter and this distinguishes phantom field from ordinary
(canonical) fields. But this distinctive property of the phantom
field causes a series of quantum vacuum instabilities. Also, due to
characteristic $p<-\rho$ feature of this kind of matter, it violates
the null energy condition. On the other hand, dark energy with
$\omega<-1$ (phantom energy) is beset with a host of undesirable
properties which makes this model of dark energy to be unattractive.
So, cosmologists attempted to find a way which removes these
problems. They are forced to consider models that realize
phantom-like behavior without introducing any phantom matter. In the
language of general relativity, phantom-like behavior means that the
equation of state parameter of dark energy is less than $-1$ or its
energy density increases with expansion of the universe. In the
cosmological models that there is no phantom matter, two mentioned
expressions for the phantom-like behavior are not equivalent and
should be investigated separately.
\section{Phantom mimicry: extended models}
\subsection{$\Lambda$CDM vs $\Lambda$DGP} \label{bozomath}
A cosmological model which has the capability to realize the
phantom-like behavior without introducing any phantom matter, is the
$\Lambda$DGP model. Firstly, we introduce briefly the $\Lambda$DGP
model. The DGP model, as a modified theory of gravity that modifies
the geometric sector of the Einstein field equations, has a modified
Friedmann equation as follows \citep{Dva00a,Def01,Lue06}
\begin{equation}
H^{2}\pm\frac{H}{r_{c}}=\frac{8\pi G}{3}\rho(t)\,,
\end{equation}
where $H(t)=\frac{\dot{a}}{a}$ is the Hubble parameter. This model,
in its self-accelerating branch (corresponding to the negative sign
of equation (2)), explains late-time accelerated expansion of the
universe. But, the normal branch (positive sign), has no
self-accelerating behavior. Nevertheless, the self-accelerating
branch has beset with ghost problem. The most trivial extended-DGP
model which can create the accelerated phase in the normal branch is
the $\Lambda$DGP model. In this model a cosmological constant which
plays the role of dark energy, lies on the brane. This scenario
based on the so-called \textit{dynamical screening} of the brane
cosmological constant, has the capacity to realize a phantom-like
behavior without introducing any phantom matter neither on the brane
nor in the bulk. In this case, brane cosmological constant as a dark
energy component is screened due to curvature modification at
late-time \citep{Sah03,Sah04,Sht00,Lue04}. The action of this model
is
\begin{equation}
{\cal{S}}=\frac{1}{2} M_{5}^{3} \int d^{5}x\sqrt{-g}{\cal{R}}+
\frac{1}{2}m_{p}^{2}\int d^{4}x\sqrt{-q}R+\int
d^{4}x\sqrt{-q}{\cal{L}}_{m}\,,
\end{equation}
where $M_{5}$ is the five-dimensional fundamental scale. The first
term in ${\cal{S}}$ is the Einstein-Hilbert action in five
dimensions for a five-dimensional metric $g_{AB}$ with Ricci scalar
${\cal{R}}$. The metric $q_{\mu\nu}$ is the induced
(four-dimensional) metric on the brane, and $q$ is its determinant.
The second integral contains an induced curvature effect that
appears in DGP setup due to quantum corrections via interaction of
the bulk graviton with matter on the brane. To apply correct
boundary conditions, the Gibbons-Hawking term containing the
extrinsic curvature of the brane should be added to the brane part
of the action. The crossover distance, that gravity in the scales
larger than it appears to be $5$-dimensional, is defined as follows
\begin{equation}
r_{c}=\frac{m_{p}^{2}}{2M_{5}^{3}}
\end{equation}
The cosmology on the brane for a spatially-flat universe follows
Deffayet's modified Friedmann equation \citep{Dva00a,Def01,Lue06}
\begin{equation}
H^{2}\pm \frac{H}{r_{c}}=\frac{8\pi
G}{3}(\rho_{m}+\rho_{\Lambda})\,.
\end{equation}
The negative and positive signs in this equation represent the
self-accelerating and normal branches respectively. In which
follows, we focus on the normal branch, because this branch has the
key property that brane is extrinsically curved so that shortcuts
through the bulk allow gravity to screen the effect of the brane
energy-momentum contents at Hubble parameter $H\sim r_{c}^{-1}$ and
there is no ghost instability in this branch
\citep{Sah03,Sah04,Sht00,Lue04}
\begin{equation}
H^{2}+\frac{H}{r_{c}}=\frac{8\pi
G}{3}\Big(\rho_{m}+\rho_{\Lambda}\Big).
\end{equation}
Comparing this equation and the standard Friedmann equation in a
spatially flat universe
\begin{equation}
H^{2}=\frac{8\pi G}{3}(\rho_{m}+\rho_{DE})\,,
\end{equation}
one can find an effective dark energy component as follow
\citep{Mel03,Rie04,Kom09}
\begin{equation}
\frac{8\pi G}{3}\rho_{DE}^{eff}=\Lambda-\frac{H}{r_{c}}.
\end{equation}
Since in this case $H(t)$ is a decreasing function of the cosmic
time, the effective dark energy component increases with time, and
therefore we realize a phantom-like behavior without introducing any
phantom matter that violets the null energy condition and suffers
from several theoretical problems. We can define a $w_{eff}$ by
using $\rho_{eff}=\rho_{0}(1+z)^{3(1+w_{eff})}$ and comprehend that
in a red-shift such as $z^{*}$ which $H(z^{*})<\Lambda r_{c}$ (
during the epoch which dark energy is dominated ), $w_{eff}$ is less
than $-1$ \citep{Laz06,Maa06,Ala06,Laz07}. In other words, in the
mentioned red-shift, two expressions of the phantom-like behavior
are equivalently applicable. Also, existence of $\frac{H}{r_{c}}$
term causes the cosmological constant $\Lambda$ to be screened, and
to be appeared less than its actual value. This case is called the
\emph{gravitational screening effect} and as we have pointed out,
one can realize phantom-like behavior based on this effect
\citep{Sah03,Sah04,Sht00,Lue04}. The $\Lambda$DGP model is an
alternative for $\Lambda CDM$ ( the cosmological constant $\Lambda$
plus the cold dark matter ). It has been shown recently that
$\Lambda$DGP in some respects gives even better fit to observations
than other dark energy scenarios, albeit with one more parameter
\citep{Sol09}. To investigate expansion history of $\Lambda$DGP and
comparing it with constant-$w$ dark energy models (such as the
$\Lambda$CDM), we study variation of luminosity distances versus the
redshift in these scenarios. By rewriting the Friedmann equation
(6) in the more phenomenological form, we find
\begin{equation}
\frac{H(z)}{H_{0}}=\sqrt{\Omega_{m}(1+z)^{3}+\Omega_{\Lambda}+\Omega_{r_{c}}}-\sqrt{\Omega_{r_{c}}}
\end{equation}
Where
\begin{equation}
\Omega_{m}=\frac{8\pi G}{3H_{0}^{2}}\rho_{0m},\ \quad
\Omega_{\Lambda}=\frac{\Lambda}{3H_{0}^{2}},\ \quad
\Omega_{r_{c}}=\frac{1}{4r_{c}^{2}H_{0}^{2}}\,.
\end{equation}
Also, by using this dimensionless equation, one can imply a
constrain on the parameters of the model as follows
\begin{equation}
\Omega_{m}+\Omega_{\Lambda}=1+2\sqrt{\Omega_{r_{c}}}\,.
\end{equation}
This constraint means that a flat $\Lambda DGP$ model mimics a
closed $\Lambda CDM$ model in the $(\Omega_{m},\Omega_{\Lambda})$
plane \citep{Laz06,Maa06,Ala06,Laz07}. Friedmann equation in the
$\Lambda CDM$ model ( constant-$w$ model with $w=-1$ ) is
phenomenologically written as follows
\begin{equation}
\frac{H(z)}{H_{0}}=\sqrt{\Omega_{m}(1+z)^{3}+(1-\Omega_{m})(1+z)^{3(1+w)}}.
\end{equation}
By considering luminosity-distance in the standard form in spatially
flat cosmology
\begin{equation}
d_{L}(z)=(1+z)\int_{0}^{z} \frac{H_{0}}{H(z)}dz \,,
\end{equation}
we can compare $d_{L}(z)$ calculated in $\Lambda$DGP and
$\Lambda$CDM frameworks. This comparison is shown in figure $3$. Up
to parameters values used to plot this figure, the luminosity
distance calculated in the $\Lambda DGP$ is larger than that of the
$\Lambda CDM$ for a given redshift.
\section{$f(R)$-DGP scenario}
As an extension of the previous subsection, here we consider
possible modification of the induced gravity on the brane in the
spirit of $f(R)$ theories
\citep{Noz08a,Ata08,Saa09,Noz09a,Noz09b,Noz09d,Noz09c,Bou09,Ata06,Ata07,Ata09}.
We assume that induced gravity on the brane can be modified by a
general $f(R)$ term. It has been shown that $4D$ $f(R)$ theories can
follow closely the expansion history of the $\Lambda$CDM universe
\citep{Hu07,Mar09}. Here we study an extension of $f(R)$ theories to
a DGP braneworld setup. By focusing on the luminosity
distance-redshift relation, we compare expansion history of this
type of model with other alternative scenarios. The action of this
model can be written as follows
\begin{equation}
{\cal{S}}=\frac{M_{5}^{3}}{2}\int d^{5}x\sqrt{-g} {\cal{R}} + \int
d^{4}x\sqrt{-q}\Big(M_{5}^{3}\overline{K}+{\cal{L}}\Big)\,,
\end{equation}
where by definition
\begin{equation}
{\cal{L}}=\frac{m_{p}^{2}}{2}f(R)+{\cal{L}}_{m}\,.
\end{equation}
By calculating the bulk-brane Einstein's equations and using a
spatially flat FRW line element, the following modified Friedmann
equation is obtained
\citep{Noz09a,Noz09b,Noz09c,Noz09d,Ata06,Ata07,Ata09,Bou09}
\begin{equation}
H^{2}=\frac{8\pi
G}{3}(\rho^{m}+\rho^{(curv)})\pm\frac{H}{\bar{r}_{c}}
\end{equation}
where
\begin{equation}
\rho^{(curv)}=m_{p}^{2}\bigg(\frac{1}{2} \Big[f(R)-R f'(R)\Big]
-3\dot{R}H f''(R) \bigg)\,,
\end{equation}
is energy density corresponding to the curvature part of the theory.
This energy density can be dubbed as \emph{dark curvature} energy
density. $\bar{r}_{c}$ is the re-scaled crossover distance that is
defined as $\bar{r}_{c}=r_{c}f'(R)$ and a prime marks
differentiation with respect to the Ricci scalar, $R$. We note that
in this scenario there is an effective gravitational constant, which
is re-scaled by $f'(R)$ so that $G=G_{eff}\equiv\frac{1}{8\pi
m_{p}^{2}f'(R)}$ \citep{Noz09a}. In order to compare this model with
other alternative scenarios, it is more suitable to rewrite the
normal branch of the Friedmann equation (16) in the following form
\begin{equation}
\frac{H(z)}{H_{0}}=\sqrt{\Omega_{m}(1+z)^3+
\Omega_{curv}(1+z)^{3(1+\omega_{curv})}+\Omega_{r_{c}}}-\sqrt{\Omega_{r_{c}}}\,,
\end{equation}
where by definition $$\Omega_{curv}=\frac{8\pi
G}{3H_{0}^{2}}\rho_{0}^{(curv)}\,\,,\quad\quad
\Omega_{r_{c}}=\frac{1}{4[r_{c}f_{0}'(R)]^{2}H_{0}^{2}}\,,$$ and
also
\begin{equation}
w_{curv}=-1+\frac{\ddot{R}f''(R)+\dot{R}\Big[\dot{R}f'''(R)-
Hf''(R)\Big]}{\frac{1}{2}[f(R)-Rf'(R)]-3H\dot{R}f''(R)}\,.
\end{equation}
We note that equation of state parameter of the curvature fluid is
not a constant; it varies actually with redshift. To proceed
further, in which follows we consider the Hu-Sawicki $f(R)$ model
\citep{Hu07,Mar09} given by
\begin{equation}
f(R)=R-m^{2}\,\frac{c_{1}(\frac{R}{m^{2}})^{n}}{c_{2}(\frac{R}{m^{2}})^{n}+1},
\end{equation}
where $m^{2}$, $c_{1}$, $c_{2}$ and $n$ are free positive parameters
that can be expressed as functions of density parameters
\citep{Hu07,Mar09}. Here we explore the dependence of these
parameters on density parameters defined in our setup. To do this
end, we follow the procedure presented in Ref. \citep{Hu07}.
Variation of the action (14) with respect to the metric yields the
induced modified Einstein equations on the brane
\begin{equation}
G_{\alpha\beta}=\frac{1}{M_{5}^{6}}{\cal{S}}_{\alpha\beta}-{\cal{E}}_{\alpha\beta}\,,
\end{equation}
where ${\cal{E}}_{\alpha\beta}$, the projection of the bulk Weyl
tensor on the brane is given by
\begin{equation}
{\cal{E}}_{\alpha\beta}=^{(5)}C_{RNS}^{M}n_{M}n^{R}g_{\alpha}^{N}g_{\beta}^{S}
\end{equation}
and ${\cal{S}}_{\alpha\beta}$ as the quadratic energy-momentum
correction into Einstein field equations is defined as follows
\begin{equation}
{\cal{S}}_{\alpha\beta}=-\frac{1}{4}
\tau_{\alpha\mu}\tau^{\mu}_{\beta}+\frac{1}{12}\tau\tau_{\alpha\beta}+
\frac{1}{8}g_{\alpha\beta}\tau_{\mu\nu}\tau^{\mu\nu}-\frac{1}{24}g_{\alpha\beta}\tau^{2}\,.
\end{equation}
$\tau_{\alpha\beta}$ as the effective energy-momentum tensor
localized on the brane is defined as \citep{Noz08a,Ata08,Saa09}
\begin{equation}
\tau_{\alpha\beta}=-m_{p}^{2}f'(R)G_{\alpha\beta}+\frac{m_{p}^{2}}{2}\Big[f(R)-R
f'(R)\Big]g_{\alpha\beta}+T_{\alpha\beta}+m_{p}^{2}\Big[\nabla_{\alpha}\nabla_{\beta}f'(R)-g_{\alpha\beta}\Box
f'(R)\Big]\,.
\end{equation}
The trace of Eq. (21), which can be interpreted as the equation of
motion for $f'(R)$\,, is obtained as
\begin{equation}
R=\frac{11}{6M_{5}^{6}}\tau^{2}\,.
\end{equation}
$\tau$, the trace of the effective energy-momentum tensor localized
on the brane is expressed as
\begin{equation}
\tau=m_{p}^{2}\Big[2f(R)-Rf'(R)-3\Box f'(R)\Big]-\rho_{m}\,,
\end{equation}
To highlight the DGP character of this generalized setup, we express
the results in terms of the DGP crossover scale defined as
$r_{c}=\frac{m_{p}^{2}}{2M_{5}^{3}}$. So the equation of motion for
$f'(R)$ is rewritten as follows
$$\frac{11}{6}r_{c}^{2}\Bigg (\Big[2f(R)-Rf'(R)\Big]^{2}+9\Big(\Box
f'(R)\Big)^{2}+6Rf'(R)\Box f'(R)-12f(R)\Box f'(R)\Bigg)$$
\begin{equation}
+\frac{11}{3}\frac{r_{c}}{M_{5}^{3}}\Big[Rf'(R)-2f(R)+3\Box
f'(R)\Big]\rho_{m}+\frac{11}{6M_{5}^{6}}\rho_{m}^{2}-R=0\,,
\end{equation}
In the next stage, we solve this equation for $\Box f'(R)$ to obtain
the following solution
\begin{equation}
\Box
f'(R)=-\Bigg[\frac{1}{2}\Big(6Rf'(R)-12f(R)\Big)+\frac{3\rho_{m}}{M_{5}^{3}\,r_{c}}\Bigg]\pm
\sqrt{\Bigg[\frac{1}{2}\Big(6Rf'(R)-12f(R)\Big)+\frac{3\rho_{m}}{M_{5}^{3}\,r_{c}}\Bigg]^{2}-\Theta}\,,
\end{equation}
where $\Theta$ is defined as
\begin{equation}
\Theta=\Bigg[\Big(2f(R)-Rf'(R)\Big)-\frac{1}{M_{5}^{3}\,r_{c}}\rho_{m}\Bigg]^{2}-\frac{6}{11r_{c}^{2}}R\,.
\end{equation}
Now we introduce an effective potential $V_{eff}$, which satisfies
the following equation
\begin{equation}
\Box f'(R)=\frac{\partial V_{eff}}{\partial f'(R)}\,,
\end{equation}
This effective potential has an extremum at $\Theta=0$\,
\begin{equation}
\Big[2f(R)-Rf'(R)\Big]-\frac{1}{M_{5}^{3}\,r_{c}}\rho_{m}=\pm\frac{1}{r_{c}}\sqrt{\frac{6}{11}R}\,,
\end{equation}
In the high-curvature regime, where $f'(R)\simeq 1$ and
$\frac{f(R)}{R}\simeq 1$\,, we recover the standard DGP result (one
can compare this result with corresponding result in Ref.
\citep{Hu07} to see the differences in this extended braneworld
scenario)
\begin{equation}
R\pm\frac{1}{r_{c}}\sqrt{\frac{6}{11}R}=\frac{2}{m_{p}^{2}}\,\rho_{m}\,.
\end{equation}
The negative and positive sign in this equation is corresponding to
the DGP self-accelerating and normal branch respectively. In which
follows, we adopt the positive sign corresponding to the normal
branch of the scenario. To investigate the expansion history of the
universe in our model, we restrict ourselves to those values of the
model parameters that yield expansion histories which are
observationally viable. We note that the Hu-Sawicki $f(R)$ function,
introduced in Ref. \citep{Hu07,Mar09}, was interpreted as a
cosmological constant in the high-curvature regime. The motivation
for that interpretation was to obtain a $\Lambda$CDM behavior in the
high curvature (in comparison with $m^{2}$) regime. Here we would
like to investigate $f(R)$ models that mimic the phantom-like
behavior on the brane in the mentioned regime. As we have pointed
out previously, the phantom-like behavior can be realized from the
dynamical screening of the brane cosmological constant. In this
respect, we apply the same strategy to our model, so that the second
term in the Hu-Sawichi $f(R)$ function ( that is, second term in the
right hand side of equation (20)) mimics the role of an effective
cosmological constant on the DGP brane. Then this term will be
screened by $\frac{H}{\overline{r}_{c}}$ term in the late time (see
the normal
branch of Eq. (16)).\\
In the case in which $R\gg m^{2}$\,, one can approximate Eq. (20) as
follows
\begin{equation}
\lim_{m^{2}/R \rightarrow 0}f(R)\approx
R-\frac{c_{1}}{c_{2}}m^{2}+\frac{c_{1}}{c_{2}^{2}}m^{2}\Big(\frac{R}{m^{2}}\Big)^{n}.
\end{equation}
During the late-time acceleration epoch, $f'_{0}(R)\simeq 1$ or
equivalently $R_{0}\gg m^{2}$ and we can apply the above
approximation. Also the curvature field is always near the minimum
of the effective potential. So, based on Eq. (31), we have
\begin{equation}
R+
\frac{1}{r_{c}}\sqrt{\frac{6}{11}R}=\frac{2}{m_{p}^{2}}\,\rho_{m}+2\frac{c_{1}}{c_{2}}m^{2}\,.
\end{equation}
Since $R$ in the $f(R)$ function is induced Ricci scalar on the
brane, we except crossover scale to affect on the constant
parameters $c_{1}$\,, $c_{2}$ and $m^{2}$. In Ref. \citep{Hu07} they
obtained $3m^{2}\equiv R_{c}=\frac{\rho_{_0m}}{m_{p}^{2}}$ that
$\rho_{_0m}$ is the present value of the matter density, But in our
setup the present value of the matter density (see Eq. (32)) is
given by
\begin{equation}
R_{c}+
\frac{\sqrt{R_{c}}}{r_{c}}\sqrt{\frac{6}{11}}=\frac{2}{m_{p}^{2}}\,\rho_{_0m}\,.
\end{equation}
If we solve this equation for $R_{c}$, we find
\begin{equation}
3m^{2}\equiv
R_{c}=1.1\Omega_{r_{c}}+6\Omega_{m}\pm\sqrt{0.55\Omega_{r_{c}}\Big(0.55\Omega_{r_{c}}+6\Omega_{m}\Big)}\,.
\end{equation}
Therefore, the DGP character of this extended modified gravity
scenario is addressed through $m^{2}$. As we have argued, at the
curvatures high compared with $m^{2}$, the second term on the right
hand side of equation (20) mimics the role of an effective
cosmological constant on the brane. In this respect, the second term
in the right hand side of equation (33) also mimics the role of a
cosmological constant on the brane in the high curvature regime.
With this motivation, we find
\begin{equation}
\frac{c_{1}}{c_{2}}\approx
\frac{18\Omega_{\Lambda}}{1.1\Omega_{r_{c}}+6\Omega_{m}\pm\sqrt{0.55\Omega_{r_{c}}\Big(0.55\Omega_{r_{c}}+6\Omega_{m}\Big)}}\,.
\end{equation}
There is also a relation for $\frac{c_{1}}{c_{2}^{2}}$ as follows
\begin{equation}
\frac{c_{1}}{c_{2}^{2}}=\frac{1-f_{0}'(R)}{n}\bigg(\frac{R_{0}}{m^{2}}\bigg)^{n+1}\,,
\end{equation}
where $\frac{R_{0}}{m^{2}}$ in our setup can be calculated as
follows: firstly, by using Eqs. (34) and (37), we find
\begin{equation} R+
\frac{\sqrt{\frac{6}{11}R}}{r_{c}}=\frac{2}{m_{p}^{2}}\,\rho_{_0m}a^{-3}+12\Omega_{\Lambda}\,,
\end{equation}
where $\rho_{_0m}$ can be omitted through Eq. (26) to obtain
\begin{equation} R+
\frac{\sqrt{\frac{6}{11}R}}{r_{c}}=\bigg(3m^{2}+\frac{m}{r_{c}}\sqrt{\frac{18}{11}}\bigg)a^{-3}+12\Omega_{\Lambda}\,.
\end{equation}
Finally, if we solve this equation for $\sqrt{R}$\,, we find the
following relation for $\frac{R_{0}}{m^{2}}$
\begin{equation}
\frac{R_{0}}{m^{2}}\approx
\Bigg(-\frac{3\Omega_{r_{c}}}{4m}+\Bigg[\Big(\frac{3\sqrt{\Omega_{r_{c}}}}{4m}\Big)^{2}+3\Big(1+
\frac{4\Omega_{\Lambda}}{m^{2}}\Big)+
\frac{5\sqrt{\Omega_{r_{c}}}}{2m}\Bigg]^{1/2}\Bigg)^{2}\,.
\end{equation}
where $m$ is given by Eq. (36). Note that we have set $H_{0}$ and
$a(t_{0})$ equal to unity. These relations tell us that the free
parameters of this model are $n$, $\Omega_{m}$, $\Omega_{r_{c}}$ and
$f_{0}'(R)$, whereas the latter one is constrained by Solar-System
tests. In fact, experimental data show that $f'(R)-1~ <10^{-6}$\,,
when $f'(R)$ is parameterized to be exactly \,$1$ \,in the far past.
To analyze the behavior of $w_{curv}$\,, we need to specify an
ansatz for the scale factor. Here we use the following form
\begin{equation}
a(t)=(t^{2}+\frac{t_{0}}{1-\nu})^{\frac{1}{1-\nu}}
\end{equation}
$\nu\neq 1$ is a free parameter \citep{Cai07}. By noting that the
Ricci scalar is $R=6(\frac{\ddot{a}}{a}+(\frac{\dot{a}}{a})^{2})$,
one can express the function $f(R)$ of equation (20) in terms of the
redshift $z$. Figure $1$ shows the variation of the effective
equation of state parameter versus the redshift. As we see in this
figure, in this class of models the \textit{curvature fluid} has an
effective phantom equation of state, $w_{curv} < -1$ at high
redshifts and then approaches the phantom divide ($w_{curv} = -1$)
at a redshift that decreases by decreasing $n$. The main point here
is that a modified induced gravity of the Hu-Sawicki type in DGP
framework, gives a phantom effective equation of state parameter for
all values of $n$. Note that all of these models reach
asymptotically to the de Sitter phase ($w_{curv}=-1$). As we will
show via a dynamical system approach, this de Sitter phase is a
stable phase.\\
Now using equation (13) we plot the luminosity distance versus the
redshift in this setup. The result is shown in figure $3$ in
comparison with other alternative models. Up to the parameter values
adopted here, the $\Lambda$DGP model has the potential to give a
better fit with $\Lambda$CDM than the $f(R)$-DGP model.
As we have mentioned previously, $w_{curv}$ varies with redshift and
is less than $-1$ which indicates that the curvature fluid plays the
role of a phantom scalar field in the equivalent scalar-tensor
theory. As we will show in the next section, in a dynamical system
approach the accelerating phase of the universe ($q<0$) for this
DGP-inspired $f(R)$ model is stable if we consider the curvature
fluid to be equivalent to a phantom scalar field ( this means that
we set $w_{curv}<-1$ ), while if we set $-1<w_{curv}<-\frac{1}{3}$
(equivalent to a quintessence scalar field ), we find that current
acceleration in the mentioned universe is a transient phenomenon.
\subsection{The phase space of $f(R)$-DGP models}
To investigate stability of the solutions presented in the previous
subsection, here we express the cosmological equations of $f(R)$-DGP
scenario in the form of an autonomous dynamical system. For this
purpose, we define the following normalized expansion variables
\begin{equation}
s=\frac{\sqrt{\Omega_{m}}}{a^{3/2}E}\,,\quad\quad
p=\frac{\sqrt{\Omega_{curv}}}{a^{3(1+w_{curv})/2}E}\,,\quad\quad
u=\frac{\sqrt{\Omega_{r_c}}}{E}\,.
\end{equation}
In this way, equation (16) with minus sign and in a dimensionless
form is written as follows
\begin{equation}
1+2u=s^{2}+p^{2}\,.
\end{equation}
This constrain means that the phase space of this scenario in the
($s$ - $p$) plane is outside of a circle with radius $1$ which is
defined as $s^{2}+p^{2}\geq 1$\,. The autonomous system is obtained
as follows
\begin{equation}
s'=\frac{3s[s^{2}+(1+2w_{curv})p^{2}-1]}{2(s^{2}+p^{2}+1)}\,,
\end{equation}
\begin{equation}
p'=\frac{3p[2s^{2}+(1+w_{curv})(p^{2}-s^{2}-1)]}{2(s^{2}+p^{2}+1)}\,.
\end{equation}
To achieve the critical points of this system one should set $s'=0$
and $p'=0$\,. Note that the critical points and their stability
depend on the value of $w_{curv}$\,. Here we investigate the
stability of critical points in two different subspaces of the model
parameter space where EoS of the curvature fluid has either a
phantom or a quintessence character. In table 1, we see that the
current accelerating phase of the universe expansion is unstable if
the curvature fluid is considered to be a quintessence scalar field
($-1<w_{curv}$) and stable if it is considered to be a phantom field
($w_{curv}<-1$) in the equivalent scalar-tensor theory. This result
is compatible with evolution of the deceleration parameter versus
the redshift as has been shown in figure $4$. When $z\rightarrow
-1$, the accelerating phase (with $q<0$) is stable for
$w_{curv}<-1$. It is necessary to mention that whenever
$w_{curv}=-1$, the phase space is 1D (here the curvature fluid plays
the role of a cosmological constant, the same as $\Lambda DGP$
model. For more details see Ref. \citep{Chi06}). Figure 2 shows the
phase space trajectories of the model.
\subsection{$F(R,\varphi)$-DGP model}
Now we extend the previous model to an even more general case that
the modified induced gravity is non-minimally coupled to a canonical
scalar field on the brane. This extension allows us to see the role
played by the non-minimal coupling of gravity and scalar degrees of
freedom in the cosmological dynamics on the brane. We consider a
general coupling between gravity and scalar degrees of freedom on
the brane \citep[see for
instance][]{Noz08a,Ata08,Saa09,Noz08b,Far00,Noz07a,Bar08,Bam09}. In
fact, inclusion of this field brings the theory to realize a smooth
crossing of the phantom divide line in a fascinating manner
\footnote{ We note that the normal branch of the pure DGP scenario (
which has the capability to describe the phantom like effect),
cannot realize crossing of the phantom divide line without
introducing a quintessence scalar field (a canonical field) on the
brane. Introduction of a quintessence field on the brane, brings the
theory to realize this interesting feature \citep{Chi06}. For
$F(R,\varphi$)-DGP scenario it is natural to expect realization of
the this feature due to wider parameter space in this case
\citep{Noz08a,Noz09a}.}.
The action of this general model is given as follows \citep{Noz09b}
\begin{equation}
S=\frac{M_{5}^{3}}{2}\int d^{5}x\sqrt{-g} {\cal{R}} + \int
d^{4}x\sqrt{-q}\Big(\frac{m_{p}^{2}}{2}
F(R,\phi)-\frac{1}{2}q^{\mu\nu}\nabla_{\mu}\phi\nabla_{\nu}
\phi-V(\phi)+M_{5}^{3}\overline{K}+{\cal{L}}_{m}\Big),
\end{equation}
where the first term shows the usual Einstein-Hilbert action in the
5D bulk. The second term on the right hand side is a generalization
of the Einstein-Hilbert action induced on the brane. This is an
extension of the scalar-tensor theories in one side and a
generalization of $f(R)$-gravity on the other side. $\overline{K}$
is the trace of the mean extrinsic curvature on the brane in the
higher dimensional bulk, corresponding to the York-Gibbons-Hawking
boundary term. We call this model as $F(R,\phi)$-DGP scenario. Note
that from the above action, the crossover scale takes the following
form \citep{Noz09a}
\begin{equation}
l_{F}=\frac{m_{p}^{2}}{2M_{5}^{3}}F'(R,\varphi)=
r_{c}\,\,F'(R,\varphi)\,,
\end{equation}
where as usual $r_{c}=\frac{m_{p}^{2}}{2M_{5}^{3}}$\,, and a prime
denotes a differentiation with respect to $R$. Since DGP scenario
accounts for embedding of the FRW cosmology at any distance scale
\citep{Noz07b}, we start with the following line-element
\begin{equation}
ds^{2}=q_{\mu\nu}dx^{\mu}dx^{\nu}+b^{2}(y,t)dy^{2}=-n^{2}(y,t)dt^{2}+
a^{2}(y,t)\gamma_{ij}dx^{i}dx^{j}+b^{2}(y,t)dy^{2}\,,
\end{equation}
where $\gamma_{ij}$ is a maximally symmetric $3$-dimensional metric
defined as $\gamma_{ij}=\delta_{ij}+k\frac{x_{i}x_{j}}{1-kr^{2}}$.
Also $k=-1,0,1$ parameterizes the spatial curvature and
$r^2=x_{i}x^{i}$. We choose the gauge $b^{2}(y,t)=1$ in the normal
Gaussian coordinates. The cosmological dynamics on the brane in this
model can be described by the following Friedmann equation
\citep[see][]{Noz08a,Ata08,Saa09,Noz09a}
\begin{equation}
H^{2}+\frac{k}{a^{2}}=\frac{1}{3m_{p}^{2}F'(R,\varphi)}\Bigg[\rho^{tot}+
\varrho\Big(1+\epsilon\sqrt{1+\frac{2}{\varrho}
\Big[\rho^{tot}-\frac{m_{p}^{2}F'(R,\varphi)}{a^{4}}{\cal{E}}_{0}\Big]}\Big)\Bigg]
\end{equation}
where $\epsilon=\pm1$ shows two different embedding of the brane in
the bulk,\, $\varrho\equiv\frac{6M_{5}^{6}}{m_{p}^{2}F'(R,\phi)}$\,
and
\,${\cal{E}}_{0}=3\Big(\frac{\dot{a}^{2}}{n^{2}}-a'^{2}+k\Big)a^{2}$\,
is a constant with respect to $y$ ( with $a'\equiv\frac{da}{dy}$),
see \citep{Dic01} for more detailed discussion on the constancy of
this quantity. Total energy density and pressure are defined as
~$\rho^{(tot)}=\rho_{m}+\rho_{\phi}+\rho^{(curv)}+\rho_{\Lambda}$\,
and\, ~$p^{(tot)}=p_{m}+p_{\phi}+p^{(curv)}+p_{\Lambda}$~
respectively. The ordinary matter on the brane has a perfect fluid
form with energy density $\rho_{m}$ and pressure $p_{m}$, while the
energy density and pressure corresponding to non-minimally coupled
quintessence scalar field and also those related to
\textit{curvature fluid} are given respectively as follows
\begin{equation}
\rho_{\varphi}=\bigg[\frac{1}{2}\dot{\phi}^{2}+n^{2}V(\phi)-
6\frac{dF}{d\phi}H\dot{\phi}\bigg]_{y=0},
\end{equation}
\begin{equation}
p_{\phi}=\bigg[\frac{1}{2n^{2}}\dot{\phi}^{2}-V(\phi)+\frac{2}{n^{2}}
\frac{dF}{d\phi}(\ddot{\phi}-\frac{\dot{n}}{n}\dot{\phi})+
4\frac{dF}{d\phi}\frac{H}{n^{2}}\dot{\phi}+\frac{2}{n^{2}}
\frac{d^{2}F}{d\phi^{2}}\dot{\phi}^{2}\bigg]_{y=0}\,,
\end{equation}
\begin{equation}
\rho^{(curv)}=m_{p}^{2}\bigg(\frac{1}{2} \bigg[F(R,\phi)-R
F'(R,\phi)\bigg] -3\dot{R}H F''(R,\phi) \bigg)\,,
\end{equation}
\begin{equation}
p^{(curv)}=m_{p}^{2}\bigg({2\dot{R}H F''(R,\phi)+\ddot{R}F''(R,\phi)
+\dot{R}^{2}F'''(R,\phi)-\frac{1}{2}\Big[ F(R,\phi)-R
F'(R,\phi)\Big]}\bigg).
\end{equation}
Also by definition $\rho_{\Lambda}\equiv\frac{\Lambda}{8\pi G}$.
Ricci scalar on the brane is given by
$$R=3\frac{k}{a^2}+\frac{1}{n^{2}}\bigg[6\frac{\ddot{a}}{a}+
6\Big(\frac{\dot{a}}{a}\Big)^{2}-6\frac{\dot{a}}{a}\frac{\dot{n}}{n}\bigg].$$
We set $n(0,t)=1$ where $y=0$ is chosen to be the location of the
brane. By neglecting the dark radiation term in equation (23), we
find
\begin{equation}
H^{2}=\frac{8\pi
G}{3}(\rho_{m}+\rho_{\varphi}+\rho^{(curv)})+\frac{\Lambda}{3}+\frac{1}{2l_{F}^{2}}+
\epsilon\sqrt{\frac{1}{4l_{F}^{4}}
+\frac{1}{2l_{F}^{2}}\Big[\frac{8\pi
G}{3}(\rho_{m}+\rho_{\varphi}+\rho^{(curv)})+\frac{\Lambda}{3}\Big]}\,,
\end{equation}
where $G=G_{eff}\equiv\frac{1}{8\pi m_{p}^{2}F'(R,\varphi)}$. By
adopting the negative sign, we have
\begin{equation}
H^{2}=\frac{8\pi
G}{3}(\rho_{m}+\rho_{\varphi}+\rho^{(curv)})+\frac{\Lambda}{3}-\frac{H}{r_{c}F'(R,\varphi)}\,.
\end{equation}
Comparing this equation with the following Friedmann equation
\begin{equation}
H^{2}=\frac{8\pi
G}{3}(\rho_{m}+\rho_{\varphi}+\rho^{(curv)})+\rho_{DE}^{eff}\,,
\end{equation}
we are led to conclude that the screening effect on the cosmological
constant is modified by $F'(R,\varphi)$ as follows
\begin{equation}
\frac{8\pi
G}{3}\rho_{DE}^{eff}=\frac{\Lambda}{3}-\frac{H}{r_{c}F'(R,\varphi)}\,.
\end{equation}
To see how this model works, we consider an explicit form of
$F(R,\varphi)$ as
$F(R,\varphi)=\frac{1}{2}(1-\xi\phi^{2})[R-(1-n)\zeta^{2}(R/\zeta^{2})^{n}]$
where $\zeta$ is a suitably chosen parameter \citep[see for
instance][]{Cap03,Sot10,Noj03,Noj04,Noj05,Noj06,Noj07,Noj08,Bam08,Car04,Ame07,Sta07}
and \citep{Cap08,Cap10}. For a spatially flat FRW geometry, the
Ricci scalar is given by $R=6\frac{\ddot{a}}{a}+
6(\frac{\dot{a}}{a})^{2}.$ One can investigate variation of the
effective dark energy density with respect to parameters $\xi$, $t$
and $n$ to see the status of phantom mimicry in this model. As has
been shown in Ref. \citep{Noz09a}, this model accounts for, and
modifies the phantom like behavior on the brane. Using the
conservation equation
$\dot{\rho}_{eff}+3H(1+\omega_{eff})\rho_{eff}=0$, fulfillment of
the condition $w_{eff}<-1$\, requires the constraint
$\frac{\dot{H}}{H}<\frac{\dot{F}'(R,\phi)}{F'(R,\phi)}$
\citep{Noz09a}. In this situation, this model has the potential to
realize the phantom-like behavior and smooth crossing of the phantom
divide line by the effective equation of state parameter. Now, in
order to compare the $F(R,\varphi)$-DGP model with $\Lambda$CDM and
also with other alternative models, we study expansion histories of
these models based on the variation of their luminosity distances
versus the redshift. The evolution of the cosmic expansion in the
normal branch of this F(R,$\varphi$)-DGP setup is given by
$$\frac{H^{2}(z)}{H_{0}^{2}}=\Omega_{m}(1+z)^{3}+\Omega_{\varphi}(1+z)^{3(1+\omega_{\varphi})}
+\Omega_{curv}(1+z)^{3(1+\omega_{curv})}+\Omega_{\Lambda}+2\Omega_{r_{c}}$$
\begin{equation}
-2\sqrt{\Omega_{r_{c}}}\sqrt{\Omega_{m}(1+z)^{3}+\Omega_{\varphi}(1+z)^{3(1+\omega_{\varphi})}
+\Omega_{curv}(1+z)^{3(1+\omega_{curv})}+\Omega_{\Lambda}+\Omega_{r_{c}}}\,\,.
\end{equation}
Note that by definition,
$\Omega_{r_{c}}=\frac{1}{4r_{c}^{2}H_{0}^{2}F'(R,\varphi)}$ and we
use the normalization $F'(R,\varphi)|_{(z=0)}\simeq 1$. This model
has the disadvantage that it contains more free parameters than
$\Lambda$CDM and even $\Lambda$DGP. However, this larger parameter
space provides, at least theoretically, more facilities to realize
some interesting features. The parameter space of this model can be
constraint at $z=0$ so that
\begin{equation}
1=\bigg(\sqrt{\Omega_{m}+\Omega_{\varphi}+\Omega_{curv}+\Omega_{\Lambda}
+\Omega_{r_{c}}}-\sqrt{\Omega_{r_{c}}}\bigg)^{2}.
\end{equation}
Since $\Omega_{r_{c}}$\, and \,
$\Omega_{m}+\Omega_{\varphi}+\Omega_{curv}+\Omega_{\Lambda}
+\Omega_{r_{c}}$ \, should be positive, there are two possibilities
for constrains on the parameters of this model as follows
\citep{Sah03,Sah04,Sht00,Lue04}
\begin{equation}
\Omega_{m}+\Omega_{\varphi}+\Omega_{curv}+\Omega_{\Lambda}
-2\sqrt{\Omega_{r_{c}}}=1.
\end{equation}
\begin{equation}
\Omega_{m}+\Omega_{\varphi}+\Omega_{curv}+\Omega_{\Lambda}
+2\sqrt{\Omega_{r_{c}}}=1.
\end{equation}
In our forthcoming analysis, we consider the first one of these
constraints. Finally, using equation (13), we compare the luminosity
distance - redshift relation of this model with other proposed
scenarios. The results are shown in figure $3$. In comparison with
$\Lambda$DGP and $f(R)$-DGP, the $F(R,\varphi)$-DGP scenario has the
potential to have a better fit with $\Lambda$CDM. The parameters
adopted in this model are $\Omega_{m}=0.27$\,,
$\Omega_{r_{c}}=0.01$\,, $\Omega_{\varphi}=0.1$\,,
$\Omega_{curv}=0.13$ and $\Omega_{\Lambda}=0.7$.\\
\subsection{$F(G,\varphi)$-DGP model}
In the $F(G,\varphi)$-DGP setup, the curvature corrections are taken
into account via incorporation of the Gauss-Bonnet invariant term in
the brane part of the action. The Gauss-Bonnet invariant should be
considered essentially in the bulk action. But, if there is a
coupling between scalar degrees of freedom on the brane and the
mentioned invariant term, then it is necessary to consider its
contribution in the field equations on the brane too. This is the
main reason for incorporating scalar field $\varphi$ in the
$F(G,\varphi)$ setup. We start with the action of this scenario as
follows \citep{Noz09,Noj07,Bam10}
\begin{equation}
{\cal{S}}=\frac{M_{5}^{3}}{2}\int d^{5}x\sqrt{-g} {\cal{R}} +\int
d^{4}x\sqrt{-q}\bigg[\frac{m_{p}^{2}}{2}R+M_{5}^{3}
\overline{K}+F(G,\phi)+{\cal{L}}_{m}\bigg].
\end{equation}
The \emph{Scalar-Gauss-Bonnet} term in this action is defined as
\citep{Noz09}
$$F(G,\phi)\equiv-\frac{1}{2}\partial_{\mu}\phi\partial^{\mu}\phi-V(\phi)+f(\phi)G(R).$$
By definition, the Gauss-Bonnet invariant $G(R)$ is given by
\begin{equation}
G(R)=R^2-4R_{\mu\nu}R^{\mu\nu}+R_{\mu\nu\alpha\beta}R^{\mu\nu\alpha\beta}.
\end{equation}
In order to discuss about cosmological dynamics on the brane, one
can achieve the generalized Friedmann equation as follows
\citep{Noz09}
\begin{equation}
H^{2}+\frac{k}{a^2}=\frac{\rho_{m}+\rho^{(GB)}}{3m_{p}^{2}}
+\frac{1}{2r_{c}^{2}}+\varepsilon\sqrt{\frac{1}{4r_{c}^{4}}+\frac{1}{r_{c}^{2}}
\Big(\frac{\rho_{m}+\rho^{GB}}{3m_{p}^{2}}\Big)}.
\end{equation}
The energy density corresponding to the Gauss-Bonnet term is defined
as
\begin{equation}
\rho^{(GB)}\equiv\frac{1}{2}\dot{\phi}^{2}+V(\phi)-24\dot{\phi}f'(\phi)H^3,
\end{equation}
and the corresponding pressure is defined as
\begin{equation}
p^{(GB)}=\frac{1}{2}\dot{\phi}^{2}-V(\phi)+8\frac{\partial}{\partial
t}\Big(H^2\dot{f}\Big)+16\dot{\phi}f'(\phi)H^3,
\end{equation}
where a dot marks differentiation with respect to the cosmic time.
One can express the Friedmann equation (65) for the normal branch of
this DGP-inspired scenario and for an spatially flat brane in a
dimensionless form as follows
$$\frac{ H^{2}(z)}{
H_{0}^{2}}=\Omega_{m}(1+z)^{3}+\Omega_{GB}(1+z)^{3(1+\omega_{GB})}+2\Omega_{r_{c}}$$
\begin{equation}
- 2\sqrt{\Omega_{r_{c}}}\sqrt{\Omega_{m}(1+z)^{3}+\Omega_{GB}(1+z)^{3(1+\omega_{GB})}+
\Omega_{r_{c}}}
\end{equation}
where by definition $\Omega_{GB}=\frac{8\pi
G}{3H_{0}^{2}}\rho_{0}^{(GB)}$. At the red-shift $z=0$, equation
(68) can be expressed as
\begin{equation}
1=\bigg(\sqrt{\Omega_{m}+\Omega_{GB}+\Omega_{r_{c}}}-\sqrt{\Omega_{r_{c}}}\bigg)^{2}.
\end{equation}
By choosing the positive sign of the square root, this constrain
equation yields
\begin{equation}
\Omega_{m}+\Omega_{GB}-2\sqrt{\Omega_{r_{c}}}=1.
\end{equation}
Note that the general relativistic limit can be recovered if we set
$\Omega_{r_{c}}=0$ (or $M_{5}=0$). In this case equation (70)
implies $\Omega_{m}+\Omega_{GB}=1$. Finally, the luminosity
distance-redshift of this model is compared with other proposed
alternatives in figure $3$.
\section{Some other probes}
To have more complete comparison between the proposed models, in
this section we study some other features of these models. The
current universe is accelerating and this can be realized by
investigating the deceleration parameter. The deceleration parameter
$q=-\frac{\ddot{a}}{\dot{a}^{2}}a$ can be expressed as
\begin{equation}
q(z)=\frac{H'(z)}{H(z)}(1+z)-1\,,
\end{equation}
where a prime marks differentiation with respect to $z$. We study
variation of $q$ versus the redshift in each of the previously
proposed models. For this purpose, we use $H(z)$ as given by
equations (9),\, (12),\, (18),\, (59) and (68). Figure $4$ compares
the deceleration parameter in each of the proposed scenarios. All of
these scenarios explain the late-time cosmic speed-up in their
normal DGP branches, but the redshift at which transition to the
accelerated phase occurs are different in these scenarios. The
$\Lambda$DGP model transits to the accelerated phase at $z\simeq0.8$
much sooner than other scenarios. The $F(R,\varphi)$-DGP model
transits to the accelerated phase at $z\simeq0.4$ much later than
other alternatives. We note however that the exact value of the
transit redshift depends on the choice of the parameters in each
model. Within the parameters values adopted here, the result are
shown in figure $4$. As another important result, we note that
late-time acceleration in the $f(R)$-DGP universe is a stable
phenomenon since we considered the curvature fluid to be a phantom
scalar field. Also the acceleration phase is a transient phenomenon
for $F(G,\varphi)$-DGP model if we set $w_{\varphi}>-1$. Our
inspection shows that cosmological dynamics in the normal branch of
$f(R)$-DGP scenario with a quintessence curvature fluid is very
similar to cosmological dynamics in the normal branch of
$F(G,\varphi)$-DGP scenario for $w_{GB}>-1$.
The next probe is the age of the universe in each of these
scenarios. The age of the universe at a given cosmological red-shift
is given as follow
\begin{equation}
t(z)=\int_{z}^{\infty} \frac{dz'}{(1+z')H(z')}\,,
\end{equation}
where $H(z)$ is given in each of the proposed scenarios. In figure
$5$ the age of the universe is compared in five alternative
scenarios. Within the parameter values adopted here, the age of the
universe in $f(R)$-DGP is larger than $\Lambda$CDM, but this age in
$F(G,\varphi)$-DGP is smaller than $\Lambda$CDM one.
\section{Summary}
In this paper, observational status of the $\Lambda$CDM\,,
$\Lambda$DGP and some extended phantom-like cosmologies as
alternatives for dark energy are studied. We compared these models
with each other by comparing variation of their luminosity distances
versus the redshift and also the age of the universe in each of
these scenarios. The common feature of the mentioned DGP-inspired
scenarios is possibility of realization of the phantom-like behavior
in their normal DGP branches. In the extended scenarios, we focused
on the role played by new ingredients such as the curvature effect,
modification of the induced gravity on the brane and even the
existence of a quintessence field non-minimally coupled to modified
induced gravity on the brane. Since all of our analysis are
preformed on the normal DGP branch of each scenario, there is no
ghost instabilities in these setups. As an example and to have an
intuition about stability of the solutions, we analyzed stability of
cosmological phases of $f(R)$-DGP scenario within a dynamical system
approach. We have analyzed the phase space of the normal branch of
this model where curvature fluid plays the role of a phantom or
quintessence scalar field in the equivalent scalar-tensor theory. We
have shown that the matter dominated phase of this model is a
repeller, unstable point. However, there is a de Sitter phase which
is an attractor point if curvature fluid plays the role of a phantom
scalar field in the equivalent scalar-tensor theory. For a
quintessence field plying the role of the curvature fluid, the de
Sitter phase is an unstable, saddle point. Up to the parameters
values adopted in this paper, $F(R,\varphi)$-DGP scenario mimics the
behavior of the $\Lambda$CDM scenario more than other proposed
models. As another important outcome, all of these scenarios explain
the late-time cosmic speed-up in their normal DGP branches, but the
redshift at which transition to the accelerating phase occurs are
different in these scenarios. For instance, the $\Lambda$DGP model
transits to the accelerating phase at $z\simeq 0.8$, much sooner
than other scenarios. However, the $F(R,\varphi)$-DGP model transits
to the accelerated phase at $z\simeq 0.4$, much later than other
alternatives. Nevertheless, the exact value of the transit redshift
depends on the choice of the parameters in each model. Finally, we
compared the age of the universe in each of the proposed scenarios.
Within the parameter values adopted in this paper, the age of the
universe in the $f(R)$-DGP model is larger than $\Lambda$CDM, but
this age in $F(G,\varphi)$-DGP is smaller than $\Lambda$CDM one.
|
\section{Introduction}
Unitary and special unitary groups play an important role in various fields of physics. Several problems that arise in the context of these groups require to express them in terms of a set of real parameters. In general, such parameterizations are not unique in the sense that the considered groups can be parameterized in many different ways. With regard to the diversity of problems it is reasonable to have a repertoire of different parameterizations available, in order to be able to choose the most convenient one for a given problem. In Ref.~\cite{Composite} we recently introduced the composite parameterization of the unitary group $\mathcal{U}(d)$ as an alternative to the canonical parameterization $U=\exp(i H)$ via hermitian matrices $H$ and those presented in Refs.~\cite{TilmaUN,Jarlskog,Dita}. It was shown that our parameterization enables a simple identification of redundant parameters when it is applied to describing orthonormal bases, density matrices of arbitrary rank and subspaces. For all these objects we found representations containing the minimal number of parameters needed. Due to its concise notation, simple implementation and computational benefits it has already found widespread applications in research on lattice correlation functions \cite{RWJohnson1}, quantum nonlocality \cite{geomNL}, genuine multipartite entanglement \cite{GenMultClass,Ma,RelClasses} and quantum secret sharing \cite{Schauer}.
The aim of the present paper is twofold. First, we adopt our concept that was used in Ref.~\cite{Composite} to obtain a novel parameterization of the special unitary group $\mathcal{SU}(d)$. The need of additional representations of this group is not only apparent because of its vital relevance in all kinds of fields involving quantum physics (see Ref.~\cite{Redkov} for an overview), but moreover is given because the number of available parameterizations is relatively low compared to $\mathcal{U}(d)$. Here, our parameterization is proposed as an alternative to the canonical parameterization $U=\exp(i H)$ via \emph{traceless} hermitian matrices $H$ and the generalized Euler angle parameterization \cite{TilmaSUN} introduced by Tilma and Sudarshan.
Second, we rigorously derive the normalized Haar measure in terms of the introduced parameters for both the unitary $\mathcal{U}(d)$ and the special unitary group $\mathcal{SU}(d)$ of arbitrary dimension. In form of an infinitesimal volume element of a group, the Haar measure contains all information about the distribution density in its parameter representation. This in return is essential for the capability of generating uniformly distributed random unitaries, density matrices and subspaces, as they are important in, for instance, Monte Carlo simulations \cite{Fresch} or quantum data hiding \cite{Terhal,Winter1}. Explicit expressions of the normalized Haar measure can furthermore be useful to tackle group integrals as they appear in lattice QCD \cite{Verbaaschot}, quantum optics \cite{Serafini}, stochastic processes \cite{Katori} and mesoscopic systems \cite{Savin}. In quantum information they can be found in the context of symmetric states (\emph{Werner} and isotropic states \cite{Werner,Eggeling,Chruscinski}) and the \emph{a priori entanglement} of quantum systems \cite{SlaterA,HaydenEnt,Gross}.
Within recent years much attention has been paid to integrals over unitary groups, i.e. $\int_{\mathcal{U}(d)} f(U,U^*) dU $ whose integrand is a polynomial in $U$ and $U^{*}$. It was shown that such integrals can be replaced by a sum of function values $f(U_i,U_i^{*})$ using a finite set of unitaries $\{ U_i \}_N$. Such sets are termed \emph{unitary $t$-designs} \cite{Gross,Dankert,Winter2}, wherein $t$ denotes that the degree of the polynomial in $U$ and $U^{*}$ is at most $t$. Whereas the existence of a unitary $t$-design was proven for all dimensions $d$ and all polynomial degrees $t$ (see Ref.~\cite{Seymour}), it is generally unknown how to construct them for arbitrary $d$ and $t$. In addition, there currently exists no simple analytic method for solving any integral over any polynomial. Due to this, there have been several attempts to find schemes to approximate unitary $t$-designs. Here, the usefulness of our results in the context of unitary $t$-designs and their approximations is obvious: Since we do not only provide the Haar measure but also the exact parameter ranges for group covering, the integration of polynomials can be performed explicitly and solved analytically in many cases. This in return allows to verify (or falsify) suggested unitary designs and to test the accuracy of approximations. Apart from that, our tools for computing integrals are not limited to polynomials, but are applicable to arbitrary functions.
The paper is organized as follows. In Section \ref{compU} we review the composite parameterization of the unitary group $\mathcal{U}(d)$. In Section \ref{compSU} we introduce the composite parameterization of the special unitary group $\mathcal{SU}(d)$. In Section \ref{haarU} the general formula for the Haar measure on $\mathcal{U}(d)$ is stated and proven. An analogous formula for the Haar measure on the special unitary group $\mathcal{SU}(d)$ can be found in Section \ref{haarSU}. Finally, in Section \ref{NotesInt} we make useful remarks on computing integrals over $\mathcal{U}(d)$ and $\mathcal{SU}(d)$ using the composite parameterization and the associated Haar measure.
\section{Composite parameterization of the unitary group $\mathcal{U}(d)$}
\label{compU}
Consider a $d$-dimensional $(d\geq2)$ complex Hilbert space $\mathcal{H}=\mathbb{C}^d$ spanned by the orthonormal basis $\{ \ket{1},\ldots,\ket{d} \}$. On this space define $d$ one-dimensional projectors
\begin{align}
\label{projector}
P_l=\ket{l}\bra{l} \hspace{4.7cm} 1 \leq l \leq d
\end{align}
and ${d(d-1)}/{2}$ anti-symmetric matrices \footnote{These can be considered as generalizations of the Pauli matrix $\sigma_y$}
\begin{align}
\label{antisymm}
Y_{m,n}=-i\ket{m} \bra{n} + i \ket{n} \bra{m} \hspace{1.2cm} 1\leq m < n \leq d
\end{align}
each acting on a two-dimensional subspace spanned by $\ket{m}$ and $\ket{n}$. In our previous paper \cite{Composite}, using these operators we have proven the following theorem:
\begin{samepage}
\begin{thm}
\label{theorem1}
Any operator of the unitary group $\mathcal{U}(d)$ can be written as \footnote{The order of the product is $\prod_{i=1}^{N}A_i=A_1 \cdot A_{2} \cdots A_N$}
\begin{align}
\label{Uc}
U_C=\left[\prod_{m=1}^{d-1} \left(\prod_{n=m+1}^{d} \exp \left( i P_n \lambda_{n,m} \right) \exp \left( i Y_{m,n} \lambda_{m,n} \right) \right) \right] \cdot \left[ \prod_{l=1}^{d} \exp(i P_l \lambda_{l,l})\right] \ ,
\end{align}
using $d^2$ real parameters $\{ \lambda_{m,n} \}_{m,n=1,\ldots,d}$ in the ranges $\lambda_{m,n} \in \left[0, 2 \pi \right]$ for $m \geq n$ and $\lambda_{m,n} \in \left[0, \frac{\pi}{2} \right]$ for $m < n$.
\end{thm}
\end{samepage}
The idea behind this construction was to compose the unitary group out of `elementary operations' such as rotations and phase shifts. Here, these operations are realized through the matrix exponential functions $\exp \left( i Y_{m,n} \alpha \right)$ (generates a rotation) and $\exp \left( i P_l \alpha \right)$ (shifts a phase). To make an ansatz for an arbitrary unitary operator it is useful to think of unitaries as orthonormal basis transformations. In this way, the form (\ref{Uc}) can be interpreted as one option of incorporating $d$ global phase operations
\begin{align}
\exp(i P_l \lambda_{l,l})
\end{align}
as well as $d(d-1)/2$ rotations followed by relative phase shifts
\begin{align}
\label{distinc}
\Lambda_{m,n}=\exp \left( i P_n \lambda_{n,m} \right) \exp \left( i Y_{m,n} \lambda_{m,n} \right)
\end{align}
that result from partitioning the Hilbert space into two-dimensional subspaces according to (\ref{antisymm}). That this $d^2$ parameter set of unitary operators indeed covers the whole unitary group $\mathcal{U}(d)$ was proven in Ref.~\cite{Composite}. This was done by showing that for any $U\in\mathcal{U}(d)$ there exists a $U_C$ such that $U_C^{\dagger} U=\mathbbm{1}$.
\section{Composite parameterization of the special unitary group $\mathcal{SU}(d)$}
\label{compSU}
Unitary operators obeying $\det U =1$ constitute a subgroup of $\mathcal{U}(d)$ called the special unitary group $\mathcal{SU}(d)$. The first new result of this paper is that this group can be parameterized similarly as $\mathcal{U}(d)$ using the concept of the composite parameterization. First, let us note that as special unitary operators satisfy an additional constraint the special unitary group $\mathcal{SU}(d)$ can be described by $d^2-1$ real parameters. As redundant parameters are undesirable we have to find a parameterization that contains exactly this number of parameters. Second, it is known that $U=\exp(i H \alpha)$ is special unitary for all $\alpha \in \mathbb{R}$ only if $H$ is hermitian and \emph{traceless}. In the composite parameterizaton of the unitary group $\mathcal{U}(d)$ we have used matrix exponentials of $P_l$ and $Y_{m,n}$ as defined in Eq.~(\ref{projector}) and Eq.~(\ref{antisymm}). The latter is already traceless but any one-dimensional projector $P_l$ has $\Tr(P_l)=1$. As the projectors $P_l$ were used to create phase shifts it is clear that they have to be replaced by a set of diagonal \emph{traceless} operators. For that we introduce the following operators
\begin{align}
\label{genZ}
Z_{m,n}=\ket{m} \bra{m} - \ket{n} \bra{n} \hspace{1.2cm} 1\leq m < n \leq d \ .
\end{align}
These are possible generalizations of the diagonal Pauli matrix $\sigma_z$ acting on the subspace spanned by $\ket{m}$ and $\ket{n}$. In (\ref{distinc}) the operation $\exp \left( i P_n \lambda_{n,m} \right)$ was used to generate a relative phase shift between the vector components $\ket{m}$ and $\ket{n}$. However, the same effect can also be achieved with $\exp \left( i Z_{m,n} \lambda_{n,m} \right)$ meaning that (\ref{distinc}) can be replaced by
\begin{align}
\label{distincSU}
\Lambda_{m,n}=\exp \left( i Z_{m,n} \lambda_{n,m} \right) \exp \left( i Y_{m,n} \lambda_{m,n} \right) \ .
\end{align}
It now remains to turn our attention to the last part of (\ref{Uc}), i.e. $\left[ \prod_{l=1}^{d} \exp(i P_l \lambda_{l,l})\right]$. Here, each $\exp(i P_l \lambda_{l,l})$ is regarded as a global phase operation on $\ket{l}$. Special unitarity implies that there are only $d-1$ independent global phase operations instead of $d$ for $\mathcal{U}(d)$. There is no unique way how these operations can be realized using matrix exponentials of (\ref{genZ}). However, a possible and convenient choice is
\begin{align}
\exp(i Z_{l,d} \lambda_{l,l}) \hspace{1.8cm} 1 \leq l \leq d-1 \ .
\end{align}
In this version, the first $d-1$ vectors $\ket{1},\ldots,\ket{d-1}$ experience the phase shifts\\$e^{i \lambda_{1,1} }\ket{1},\ldots,e^{i \lambda_{d-1,d-1} }\ket{d-1}$ while the last vector $\ket{d}$ gets phase shifted in the overall inverse direction, i.e. $e^{-i \sum_{l=1}^{d-1}\lambda_{l,l}}\ket{d}$. Note that according to our labeling there is no parameter $\lambda_{d,d}$. Thus, in total we have the desired number of $d^2-1$ parameters $\lambda_{m,n}$. In summary, this leads to the next theorem:
\begin{samepage}
\begin{thm}
\label{theorem2}
Any operator of the special unitary group $\mathcal{SU}(d)$ can be written as
\begin{align}
\label{SUc}
U_{C}=\left[\prod_{m=1}^{d-1} \left(\prod_{n=m+1}^{d} \exp \left( i Z_{m,n} \lambda_{n,m} \right) \exp \left( i Y_{m,n} \lambda_{m,n} \right) \right) \right] \cdot \left[ \prod_{l=1}^{d-1} \exp(i Z_{l,d} \lambda_{l,l})\right] \ ,
\end{align}
using $d^2-1$ real parameters $\{ \lambda_{m,n} \}$\footnote{Note that the indices $m$ and $n$ again run from $1$ to $d$ except that there is no $\lambda_{d,d}$.} in the ranges $\lambda_{m,n} \in \left[0, \pi \right]$ for $m > n$, $\lambda_{m,n} \in \left[0, \frac{\pi}{2} \right]$ for $m < n$ and $\lambda_{m,n} \in \left[0, 2\pi \right]$ for $m = n$.
\end{thm}
\end{samepage}
\begin{proof}
Proving that any $U\in \mathcal{SU}(d)$ may be written as (\ref{SUc}) is equivalent to showing that $U_C^{\dagger}U=\mathbbm{1}$ can be achieved for all group elements. Let $U=\sum_{r,s=1}^{d} a_{r,s} \ket{r}\bra{s}$ be an arbitrary special unitary operator, i.e. $U$ fulfils $\sum_{i=1}^{d} a_{m,i}^{*}a_{n,i}=\sum_{i=1}^{d} a_{i,m}^{*}a_{i,n}=\delta_{mn}$ and $\det U=1$.
The conjugate transpose of $U_C$ as given in (\ref{SUc}) using the abbreviation (\ref{distincSU}) is
\begin{align}
U_C^{\dagger}=\left[ \prod_{l=1}^{d-1} \mbox{exp}(-i Z_{d-l,d} \lambda_{d-l,d-l})\right] \cdot \left[\prod_{m=1}^{d-1} \left(\prod_{n=1}^{m} \Lambda_{d-m,d+1-n}^{\dagger} \right) \right] \ .
\end{align}
The order of the factors in $U_C^{\dagger}$ implies that $\Lambda_{1,2}^{\dagger}$ acts first on $U$. For $U'=\Lambda_{1,2}^{\dagger} U=\sum_{r,s=1}^{d} a'_{r,s} \ket{r}\bra{s}$ one obtains
\begin{align}
\label{proofa}
a'_{1,s}&=e^{-i\lambda_{2,1}}\cos(\lambda_{1,2})a_{1,s}-e^{i\lambda_{2,1}}\sin(\lambda_{1,2})a_{2,s} \ , \\
a'_{2,s}&=e^{-i\lambda_{2,1}}\sin(\lambda_{1,2})a_{1,s}+e^{i\lambda_{2,1}}\cos(\lambda_{1,2})a_{2,s} \ .
\end{align}
All other components remain unchanged, i.e. $a'_{r,s}=a_{r,s}$ for $r>2$ if $d>2$. We observe that $a'_{2,1}$ can always be made zero using particular values for $\lambda_{1,2}$ and $\lambda_{2,1}$: In case $a_{1,1}$ and $a_{2,1}$ both are zero, both parameters $\lambda_{1,2}$ and $\lambda_{2,1}$ can be chosen freely. If only $a_{1,1}=0$ one chooses $\lambda_{1,2}=\frac{\pi}{2}$. If both are unequal zero then $a'_{2,1}$ vanishes for
\begin{align}
\arg(e^{i2\lambda_{2,1}}a_{2,1})&=\arg(-a_{1,1}) \ ,\\
\tan(\lambda_{1,2})&=\frac{|a_{2,1}|}{|a_{1,1}|} \ .
\label{proofb}
\end{align}
This can always be achieved with $\lambda_{2,1}\in [0,\pi]$ and $\lambda_{1,2}\in [0,\frac{\pi}{2}]$. Analogously, one can make the component $a''_{3,1}$ of $U''=\Lambda^{\dagger}_{1,3}U'$ zero for the case $d>2$. In this way, all components $\overline{a}_{r,s}$ with $r>s$ of $\overline{U}=\prod_{m=1}^{d-1} \left(\prod_{n=1}^{m} \Lambda_{d-m,d+1-n}^{\dagger} \right) U=\sum_{r,s=1}^{d} \overline{a}_{r,s} \ket{r}\bra{s}$ can be made zero. The unitarity constraints $\sum_{i=1}^{d} \overline{a}_{m,i}^{*}\overline{a}_{n,i}=\sum_{i=1}^{d} \overline{a}_{i,m}^{*}\overline{a}_{i,n}=\delta_{mn}$ imply that also all $\overline{a}_{r,s}$ with $r<s$ have become zero during this procedure. Hence, $\overline{U}$ is diagonal $\overline{U}=\sum_{r=1}^{d} \overline{a}_{r,r} \ket{r}\bra{r}$ where $\overline{a}_{r,r}$ are complex numbers of magnitude one, i.e. $\overline{a}_{r,r}=e^{i\alpha_r}$. The first $d-1$ phases $\alpha_1,\ldots,\alpha_{d-1}$ can be compensated via $\left[ \prod_{l=1}^{d-1} \mbox{exp}(-i Z_{d-l,d} \lambda_{d-l,d-l})\right] \overline{U}$ by choosing $\lambda_{r,r}=\alpha_r$. This is guaranteed to be achievable with $\lambda_{r,r}\in [0, 2\pi ]$. Recall that so far we have only multiplied special unitary operators implying that $U_C^{\dagger}U=\left[ \prod_{l=1}^{d-1} \mbox{exp}(-i Z_{d-l,d} \lambda_{d-l,d-l})\right] \overline{U}$ is still a member of $\mathcal{SU}(d)$. We have achieved that $U_C^{\dagger}U$ is diagonal and that the first $d-1$ diagonal entries are all equal $1$. Now, the fact that $\det(U_C^{\dagger}U) =1$ still holds implies that also the last diagonal entry equals $1$. Hence, $U_C^{\dagger}U=\mathbbm{1}$, which proves the theorem.
\end{proof}
\subsection{Remarks on the composite parameterzation}
\label{remarksUSU}
It is worth mentioning some properties of the composite parameterization. In our previous paper \cite{Composite} on the parameterization of $\mathcal{U}(d)$ we have shown that it can be rather insightful to gather the parameters $\lambda_{m,n}$ in a matrix
\begin{align}
\begin{array}{c}
\ \\
\ \\
\mbox{relative phases} \rightarrow
\end{array}
\left[
\begin{array}{ccc}
\lambda_{1,1} & \cdots & \lambda_{1,d} \\
\vdots & \ddots & \vdots \\
\lambda_{d,1} & \cdots & \lambda_{d,d} \\
\end{array}
\right]
\begin{array}{c}
\leftarrow \mbox{rotations} \\
\ \\
\
\end{array} \ .
\end{align}
In this notation the lower left entries represent relative phase shifts, the diagonal global phase shifts and the upper right rotations (with respect to the computational basis $\{ \ket{1},\ldots,\ket{d} \}$). Using this representation is particularly useful for illustrating which parameters are irrelevant for certain tasks. For instance, we have shown that for parameterizing an orthonormal set of $k$ vectors $\{ \ket{\Psi_1}, \ldots, \ket{\Psi_k} \}$ only the $k(2d-k-1)$ non-zero parameters of the following matrix are relevant
\begin{align}
& \left[
\begin{array}{ccccccc}
0 & \lambda_{1,2} & \cdots & \lambda_{1,k+1} & \cdots & \lambda_{1,d} \\
\lambda_{2,1} & \ddots & \ddots & \vdots & \ddots & \vdots \\
\vdots & \ddots& 0 & \lambda_{k,k+1} & \cdots & \lambda_{k,d} \\
\lambda_{k+1,1} & \cdots & \lambda_{k+1,k} & 0 & \cdots & 0 \\
\vdots & \ddots & \vdots & \vdots & \ddots & \vdots\\
\lambda_{d,1} & \cdots & \lambda_{d,k} & 0 & \cdots & 0 \\
\end{array}
\right]
\begin{array}{c}
\left. {\begin{array}{c}
\\
\\
\\
\end{array}} \right\} k \hspace{7mm} \\
\\
\left. {\begin{array}{c}
\\
\\
\\
\end{array}} \right\} d-k
\end{array} \\
& \hspace{0.6cm} \underbrace{ \hspace{3.1cm} }_{k} \hspace{7mm} \underbrace{ \hspace{2.8cm} }_{d-k} \hspace{2cm} . \nonumber
\end{align}
A similar example was given for parameterizing the set of $k$-dimensional subspaces in $\mathbb{C}^d$ where the corresponding $2k(d-k)$ relevant parameters $\lambda_{m,n}$ were illustrated. For the composite parameterization of the special unitary group $\mathcal{SU}(d)$ all this remains valid as the operations (\ref{distinc}) and (\ref{distincSU}) are equivalent up to a global phase and are applied in the same order in both cases. Using this matrix representation for the special unitary group one should only keep in mind that there is no diagonal element $\lambda_{d,d}$.
Let us illustrate another important feature of the composite parameterization. The $(d-k)^2$ or $(d-k)^2-1$, respectively, non-zero parameters
\begin{align}
& \left[
\begin{array}{ccccccc}
0 & 0 & \cdots & 0 & \cdots & 0 \\
0 & \ddots & \ddots & \vdots & \ddots & \vdots \\
\vdots & \ddots& 0 & 0 & \cdots & 0 \\
0 & \cdots & 0 & \lambda_{k+1,k+1} & \cdots & \lambda_{k+1,d} \\
\vdots & \ddots & \vdots & \vdots & \ddots & \vdots\\
0 & \cdots & 0 & \lambda_{d,k+1} & \cdots & \lambda_{d,d} \\
\end{array}
\right]
\begin{array}{c}
\left. {\begin{array}{c}
\\
\\
\\
\end{array}} \right\} k \hspace{7mm} \\
\\
\left. {\begin{array}{c}
\\
\\
\\
\end{array}} \right\} d-k
\end{array} \\
& \hspace{0.6cm} \underbrace{ \hspace{1.9cm} }_{k} \hspace{8mm} \underbrace{ \hspace{3.1cm} }_{d-k} \hspace{2cm} \nonumber
\end{align}
correspond to the (special) unitary group for the $(d-k)$-dimensional subspace defined by $\spann(\ket{k+1},\ldots,\ket{d})$. This directly follows from the correct number of required parameters in combination with the fact that the subspace $\spann(U_C\ket{1},\ldots,U_C\ket{k})$ is independent of the illustrated parameters. Note that this feature will be helpful in an upcoming proof.
\section{Haar measure on the unitary group $\mathcal{U}(d)$}
\label{haarU}
Let us now assign an infinitesimal volume element $dU_d$ to the unitary group $\mathcal{U}(d)$ in terms of the composite parameterization $U_C=U_C(\lambda_{1,1},\ldots,\lambda_{d,d})$. This can be achieved by determining the associated Haar measure. This means that, as for any compact Lie group, we must find a measure of volume which is left and right invariant \cite{Haarorg,Folland}. Explicitly, we require that $dU_d$ satisfies
\begin{align}
dU_d=d(U_C) = d(U_1 U_C) = d(U_C U_2) \ ,
\end{align}
for all $U_1, U_2 \in \mathcal{U}(d)$. Generally, an invariant measure is (up to an irrelevant constant) determined by the absolute value of the Jacobian determinant
\begin{align}
J_d= \left| \det (j_{k,l}) \right|=\left| \det \frac{\partial(u_1,\ldots,u_{d^2})}{\partial (\alpha_1,\ldots,\alpha_{d^2})} \right| \ ,
\end{align}
wherein $\{ u_k \}$ are coefficients of the unitary $U=U(\alpha_1,\ldots,\alpha_{d^2})$ expanded in an orthogonal operator basis $\{ b_k \}$ of $\mathbb{C}^d \times \mathbb{C}^d$, i.e.
\begin{align}
u_k=\frac{\Tr(b_k^{\dagger}U)}{\Tr(b_k^{\dagger}b_k)} \hspace{1.8cm} \left( \Rightarrow U=\sum_{k=1}^{d^2} u_k b_k \right) \ ,
\end{align}
and where $\{ \alpha_l \}$ are any $d^2$ parameters that cover $\mathcal{U}(d)$. Here, a left or right translation $U_1, U_2 \in \mathcal{U}(d)$ merely induces a unitary basis transformation of $\{ b_k \}$ which is length and angle preserving. Hence, $J_d$ is invariant under these transformations and
\begin{align}
dU_d=J_d \prod_{l=1}^{d^2} d\alpha_l
\end{align}
is a Haar measure\footnote{Intuitively: When changing from the orthogonal coordinates $\{ u_k \}$ to the non-orthogonal coordinates $\{ \alpha_l \}$ the infinitesimal volume element transforms as $\prod_{k=1}^{d^2}du_k=\left| \det \frac{\partial(u_1,\ldots,u_{d^2})}{\partial (\alpha_1,\ldots,\alpha_{d^2})} \right| \prod_{l=1}^{d^2} d\alpha_l$ according to the Jacobian determinant.}. Our aim is to derive a general expression of $dU_d$ for arbitrary $d$ in terms of the parameterization introduced in Section \ref{compU}, i.e.
\begin{align}
\label{Haargeneral}
dU_d=\frac{J_d}{N_d} \prod_{k,l=1}^{d}d\lambda_{k,l} \ ,
\end{align}
where $N_d$ is a normalization constant such that $\int_{\mathcal{U}(d)} dU_d =1$. We obtain the following:
\begin{samepage}
\begin{thm}
\label{theorem3}
In terms of the $d^2$ parameters $\lambda_{m,n}$ introduced in Theorem \ref{theorem1} the (normalized) Haar measure on the unitary group $\mathcal{U}(d)$ reads
\begin{align}
\label{HaarU}
dU_d=\frac{1}{N_d}\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \prod_{k,l=1}^{d}d\lambda_{k,l} \ ,
\end{align}
with
\begin{align}
\label{normiconstanti}
N_d=\frac{(2\pi)^{d(d+1)/2}}{\prod_{m=1}^{d-1}\prod_{n=m+1}^{d} 2(n-m)}
\end{align}
such that $\int_{\mathcal{U}(d)}dU_d=1$.
\end{thm}
\end{samepage}
\begin{proof}
For simplicity, let us start with the case $d=2$. Using the canonical operator basis $\{b_k\}=\{ \ket{1}\bra{1},\ket{1}\bra{2},\ket{2}\bra{1},\ket{2}\bra{2} \}$ and the order of the parameters $\{\alpha_l\}=\{\lambda_{1,1},\lambda_{1,2},\lambda_{2,1},\lambda_{2,2}\}$ we obtain the Jacobian matrix
\begin{align}
\label{jacobianU2}
&\frac{\partial(u_1,u_2,u_3,u_4)}{\partial (\lambda_{1,1},\lambda_{1,2},\lambda_{2,1},\lambda_{2,2})}=\\
&\left(
\begin{array}{cccc}
i e^{i \lambda_{1,1}} \cos \lambda_{1,2} & -e^{i\lambda_{1,1}} \sin \lambda_{1,2} & 0 & 0 \\
0 & e^{i \lambda_{2,2}} \cos \lambda_{1,2} & 0 & i e^{i \lambda_{2,2}} \sin \lambda_{1,2} \\
-i e^{i\lambda_{1,1}+i\lambda_{2,1}} \sin \lambda_{1,2} & -e^{i\lambda_{1,1}+i\lambda_{2,1}} \cos
\lambda_{1,2} & -i e^{i\lambda_{1,1}+i\lambda_{2,1}} \sin \lambda_{1,2} & 0 \\
0 & -e^{i\lambda_{2,1}+i\lambda_{2,2}} \sin \lambda_{1,2} & i e^{i\lambda_{2,1}+i\lambda_{2,2}} \cos
\lambda_{1,2} & i e^{i\lambda_{2,1}+i\lambda_{2,2}} \cos \lambda_{1,2}
\end{array}
\right) . \nonumber
\end{align}
Using the Laplace expansion and elementary simplifications one finds
\begin{align}
J_2=2\sin(\lambda_{1,2})\cos(\lambda_{1,2})=\sin(2\lambda_{1,2}) \ .
\end{align}
The relation (\ref{Haargeneral}) combined with normalization
\begin{align}
\int_{\lambda_{2,2}=0}^{2\pi}\int_{\lambda_{2,1}=0}^{2\pi}\int_{\lambda_{1,2}=0}^{\pi/2}\int_{\lambda_{1,1}=0}^{2\pi} \frac{J_2}{N_2} d\lambda_{1,1}d\lambda_{1,2}d\lambda_{2,1}d\lambda_{2,2}=1 \ ,
\end{align}
yields the normalized Haar measure
\begin{align}
\label{U2}
dU_2=\frac{1}{4\pi^3}\sin(\lambda_{1,2})\cos(\lambda_{1,2}) d\lambda_{1,1}d\lambda_{1,2}d\lambda_{2,1}d\lambda_{2,2}\ ,
\end{align}
which is in accordance with Theorem \ref{theorem3}.
One could in principle compute $dU_d$ analogously for arbitrary $d$, i.e. using the canonical operator basis $\{b_k\}=\{\ket{1}\bra{1},\ket{1}\bra{2},\ket{1}\bra{3},\ldots,\ket{d}\bra{d-1},\ket{d}\bra{d}\}$ and the naive order $\{\alpha_l\}=\{\lambda_{1,1}, \lambda_{1,2}, \lambda_{1,3},\ldots, \lambda_{d,d-1}, \lambda_{d,d}\}$. For instance, a long and cumbersome computation reveals that
\begin{align}
\label{U3}
dU_3=\frac{1}{4\pi^6}\sin(\lambda_{1,2})\cos(\lambda_{1,2})\sin(\lambda_{1,3})\cos^3(\lambda_{1,3})\sin(\lambda_{2,3})\cos(\lambda_{2,3})\prod_{k,l=1}^{3}d\lambda_{k,l}
\end{align}
demonstrating the validity of Theorem \ref{theorem3} for $d=3$. Unfortunately, in this way, computing the determinant of the $d^2 \times d^2$ Jacobian matrix becomes increasingly unfeasible the larger $d$ gets. More importantly, this approach is not suitable to prove a general expression such as (\ref{HaarU}) for all $d$. However, the Jacobian matrix can be considerably simplified by taking into account the invariance of the Haar measure, the structure of the composite parameterization as well as the freedom in the choice of the operator basis and the order of the derivatives $\partial / \partial \lambda_{x,y}$. In this way, the correctness of Theorem \ref{theorem3} can be verified for all $d$.
First, due to the left invariance of the Jacobian determinant $J_d=\left| \det (j_{k,l}) \right|$ we are allowed to perform any transformation
\begin{align}
j_{k,l}=\frac{\partial u_k}{\partial \alpha_l}=\frac{\Tr(b_k^{\dagger} \frac{\partial U}{\partial \alpha_l})}{\Tr(b_k^{\dagger}b_k)} \hspace{0.5cm} \longrightarrow \hspace{0.5cm} j'_{k,l}=\frac{\Tr(b_k^{\dagger} U_1 \frac{\partial U}{\partial \alpha_l})}{\Tr(b_k^{\dagger}b_k)} \ ,
\end{align}
with $U_1 \in \mathcal{U}(d)$. Here, it is beneficial to choose $U_1=-iU_C^{\dagger}$ since the matrix $-iU_C^{\dagger} \frac{\partial U_C}{\partial \lambda_{x,y}}$ has a simpler form due to the fact that $U_C^{\dagger}$ and $\frac{\partial U_C}{\partial \lambda_{x,y}}$ cancel each other out partially. For instance, since all projectors $P_l=\ket{l}\bra{l}$ commute, for any derivative with respect to a global phase transformation $\partial / \partial \lambda_{l,l}$ one obtains
\begin{align}
\label{diagglobal}
-iU_C^{\dagger} \frac{\partial U_C}{\partial \lambda_{l,l}}
=-iU_C^{\dagger}U_C i P_l
=P_l=\ket{l}\bra{l} \ .
\end{align}
As can directly be inferred from the structure of $U_C$ (\ref{Uc}), the derivatives $\partial U_C / \partial \lambda_{x,y}$ are
\begin{align*}
\left[\prod_{m=1}^{x-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \left[\prod_{n=x+1}^{y} \Lambda_{x,n} \right] iY_{x,y}\left[\prod_{n=y+1}^{d} \Lambda_{x,n} \right]\left[\prod_{m=x+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=1}^{d} \exp(i P_l \lambda_{l,l})
\end{align*}
for $x<y$, and
\begin{align*}
\left[\prod_{m=1}^{y-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \left[\prod_{n=y+1}^{x-1} \Lambda_{y,n} \right] iP_{x}\left[\prod_{n=x}^{d} \Lambda_{y,n} \right]\left[\prod_{m=y+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=1}^{d} \exp(i P_l \lambda_{l,l})
\end{align*}
for $x>y$. Consequently, if we apply $-iU_C^{\dagger}$ from the left, the products to the left of $iY_{x,y}$ (respectively $iP_{x}$) cancel out and we get
\begin{align}
\label{derivs}
-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}\, = \,\left\{\begin{array}{ll} U_{x,y}^{\dagger}Y_{x,y}U_{x,y} \hspace{1.3cm} & \mbox{for} \ \ x<y \ , \\
U_{x,y}^{\dagger}P_{x}U_{x,y} & \mbox{for} \ \ x>y \ , \end{array}\right.
\end{align}
where
\begin{align}
\label{drotU}
U_{x,y}\, = \,\left\{\begin{array}{ll} \left[\prod_{n=y+1}^{d} \Lambda_{x,n} \right]\left[\prod_{m=x+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=x}^{d} \exp(i P_l \lambda_{l,l}) & \mbox{for} \ \ x<y \ , \\
\left[\prod_{n=x}^{d} \Lambda_{y,n} \right]\left[\prod_{m=y+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=y}^{d} \exp(i P_l \lambda_{l,l}) & \mbox{for} \ \ x>y \ . \end{array}\right.
\end{align}
Here, it is important to realize that $U_{x,y}^{\dagger}Y_{x,y}U_{x,y}$ and $U_{x,y}^{\dagger}P_{x}U_{x,y}$ do no longer contain operations $\Lambda_{m,n}$ with $m< \min\{x,y\}$, meaning that there are no off-diagonal elements $\ket{m}\bra{n}$ and $\ket{n}\bra{m}$ with $m< \min\{x,y\}$. This and the observation (\ref{diagglobal}) imply that when we choose the following orthogonal operator basis and order
\begin{align}
\label{basisorder}
\begin{array}{cccccccc}
b_1&=&\ket{1}\bra{2}+\ket{2}\bra{1}\ ,& \hspace{1cm} &\alpha_1&=&\lambda_{2,1}\ ,\\
b_2&=&-i\ket{1}\bra{2}+i\ket{2}\bra{1}\ ,& \hspace{1cm} &\alpha_2&=&\lambda_{1,2}\ ,\\
b_3&=&\ket{1}\bra{3}+\ket{3}\bra{1}\ ,& \hspace{1cm} &\alpha_3&=&\lambda_{3,1}\ ,\\
b_4&=&-i\ket{1}\bra{3}+i\ket{3}\bra{1}\ ,& \hspace{1cm} &\alpha_4&=&\lambda_{1,3}\ ,\\
&\vdots& & & &\vdots\\
b_{2(d-1)-1}&=&\ket{1}\bra{d}+\ket{d}\bra{1}\ ,& \hspace{1cm} &\alpha_{2(d-1)-1}&=&\lambda_{d,1}\ ,\\
b_{2(d-1)}&=&-i\ket{1}\bra{d}+i\ket{d}\bra{1}\ ,& \hspace{1cm} &\alpha_{2(d-1)}&=&\lambda_{1,d}\ ,\\
b_{2(d-1)+1}&=&\ket{2}\bra{3}+\ket{3}\bra{2}\ ,& \hspace{1cm} &\alpha_{2(d-1)+1}&=&\lambda_{3,2}\ ,\\
b_{2(d-1)+2}&=&-i\ket{2}\bra{3}+i\ket{3}\bra{2}\ ,& \hspace{1cm} &\alpha_{2(d-1)+2}&=&\lambda_{2,3}\ ,\\
b_{2(d-1)+3}&=&\ket{2}\bra{4}+\ket{4}\bra{2}\ ,& \hspace{1cm} &\alpha_{2(d-1)+3}&=&\lambda_{4,2}\ ,\\
b_{2(d-1)+4}&=&-i\ket{2}\bra{4}+i\ket{4}\bra{2}\ ,& \hspace{1cm} &\alpha_{2(d-1)+4}&=&\lambda_{2,4}\ ,\\
&\vdots& & & &\vdots\\
b_{d^2-d-1}&=&\ket{d-1}\bra{d}+\ket{d}\bra{d-1}\ ,& \hspace{1cm} &\alpha_{d^2-d-1}&=&\lambda_{d,d-1}\ ,\\
b_{d^2-d}&=&-i\ket{d-1}\bra{d}+i\ket{d}\bra{d-1}\ ,& \hspace{1cm} &\alpha_{d^2-d}&=&\lambda_{d-1,d}\ ,\\
b_{d^2-d+1}&=&\ket{1}\bra{1}\ ,& \hspace{1cm} &\alpha_{d^2-d+1}&=&\lambda_{1,1}\ ,\\
b_{d^2-d+2}&=&\ket{2}\bra{2}\ ,& \hspace{1cm} &\alpha_{d^2-d+2}&=&\lambda_{2,2}\ ,\\
&\vdots& & & &\vdots\\
b_{d^2}&=&\ket{d}\bra{d}\ ,& \hspace{1cm} &\alpha_{d^2}&=&\lambda_{d,d}\ ,&\\
\end{array}
\end{align}
the Jacobian matrix becomes a lower block-triangular matrix\\
\setlength{\unitlength}{1cm}
\begin{picture}(14,9)(0,0)
\put(3,4.5){\fontsize{12}{12}\selectfont\makebox(0,0)[]{$(j'_{k,l}) \ = $ \strut}}
\put(5,0){\includegraphics[scale=0.4]{jacobian.eps}}
\end{picture}\\
where all entries outside the grey-shaded blocks are zero. Thus, the Jacobian determinant simplifies to a product of the determinants of the blocks $M_i$ and $D$
\begin{align}
J_d=\left| \det (j'_{k,l}) \right|= \left(\prod_{i=1}^{d-1}\left| \det M_i \right| \right) \left| \det D \right| \ .
\end{align}
Furthermore, we have that $\left| \det D \right|=1$ since $D$ is a $d\times d$ identity matrix due to (\ref{diagglobal}) and the choice (\ref{basisorder}). It now remains to investigate the blocks $M_i$. Let us first consider $-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}$ for the Hilbert space $\mathcal{H}'=\mathbb{C}^{d-1}$ whose dimension is lower by one. In this case we have
\begin{align}
\label{discussion}
-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}\, = \,\left\{\begin{array}{ll} U_{x,y}^{\dagger}Y_{x,y}U_{x,y} \hspace{1.3cm} & \mbox{for} \ \ x<y \\
U_{x,y}^{\dagger}P_{x}U_{x,y} & \mbox{for} \ \ x>y \ , \end{array}\right.
\end{align}
where
\begin{align}
U_{x,y}\, = \,\left\{\begin{array}{ll} \left[\prod_{n=y+1}^{d-1} \Lambda_{x,n} \right]\left[\prod_{m=x+1}^{d-2} \prod_{n=m+1}^{d-1} \Lambda_{m,n} \right] \prod_{l=x}^{d-1} \exp(i P_l \lambda_{l,l}) \hspace{0.3cm} & \mbox{for} \ \ x<y \\
\left[\prod_{n=x}^{d-1} \Lambda_{y,n} \right]\left[\prod_{m=y+1}^{d-2} \prod_{n=m+1}^{d-1} \Lambda_{m,n} \right] \prod_{l=y}^{d-1} \exp(i P_l \lambda_{l,l}) & \mbox{for} \ \ x>y \ . \end{array}\right.
\end{align}
By substituting $\lambda_{m,n} \rightarrow \lambda_{m+1,n+1}$, $\ket{m} \rightarrow \ket{m+1}$ and $\bra{m} \rightarrow \bra{m+1}$ one can directly see that these matrices are equal to $-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}$ with $x,y\geq2$ and $U_C$ for the Hilbert space $\mathcal{H}=\mathbb{C}^{d}$ of full dimension $d$. Hence, since $J_d=\left| \det M_1 \right|\prod_{i=2}^{d-1}\left| \det M_i \right|$ we have established the recursion formula
\begin{align}
\label{recursion}
J_d=\left| \det M_1 \right|J_{d-1} \ ,
\end{align}
where the parameters in $J_{d-1}$ are to be substituted according to $\lambda_{m,n}\rightarrow\lambda_{m+1,n+1}$. This relation is a direct consequence of the fact that each $-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}$ with $x,y\geq 2$ only contains parameters $\lambda_{m,n}$ with $m,n\geq 2$. As discussed in Section \ref{remarksUSU} the set of $(d-1)^2$ parameters $\lambda_{m,n}$ satisfying $m,n\geq 2$ correspond to the unitary group $\mathcal{U}(d-1)$ for the $(d-1)$-dimensional subspace $\spann(\ket{2},\ldots,\ket{d})$. Consequently, (\ref{recursion}) simply reflects the well-known fact \cite{TilmaSUN,TilmaUN} that the infinitesimal volume element $dU_d$ on $\mathcal{U}(d)$ is the product of the infinitesimal volume element $dU_{d-1}$ on $\mathcal{U}(d-1)$ times a to-be-determined function $g(d)$, i.e.
\begin{align}
\label{invrel}
dU_d=g(d)dU_{d-1} \ .
\end{align}
Now, due to the above mentioned reasons, for the composite parameterization $dU_{d-1}$ is a function of the parameters $\lambda_{m,n}$ with $m,n\geq 2$. In addition to that, g(d) is a function of the $2d-1$ parameters $\{\lambda_{1,1},\lambda_{1,2},\ldots,\lambda_{1,d}\}$ and $\{\lambda_{2,1},\ldots,\lambda_{d,1}\}$ since they extend the unitary group $\mathcal{U}(d-1)$ acting on $\spann(\ket{2},\ldots,\ket{d})$ to the entire $\mathcal{U}(d)$ on $\mathcal{H}=\mathbb{C}^{d}$, hence, $g(d)=g(d;\lambda_{1,1},\lambda_{1,2},\ldots,\lambda_{1,d};\lambda_{2,1},\ldots,\lambda_{d,1})$. For the sake of clarity, let us illustrate this by the example $dU_2 \rightarrow dU_3$ by comparing (\ref{U2}) and (\ref{U3}): $dU_2$ is (up to a constant) given by $\sin\lambda_{1,2} \cos\lambda_{1,2} d\lambda_{1,1}d\lambda_{1,2}d\lambda_{2,1}d\lambda_{2,2}$. According to our considerations $dU_3$ must contain the same expression in terms of $\lambda_{2,2},\lambda_{2,3},\lambda_{3,2},\lambda_{3,3}$ and it must look like
\begin{align}
dU_3\sim g(3;\lambda_{1,1},\lambda_{1,2},\lambda_{1,3};\lambda_{2,1},\lambda_{3,1})\sin \lambda_{2,3} \cos \lambda_{2,3} d\lambda_{2,2}d\lambda_{2,3}d\lambda_{3,2}d\lambda_{3,3} \ .
\end{align}
Equation (\ref{U3}) shows that this is indeed the case since it can easily be observed that
\begin{align*}
g(3;\lambda_{1,1},\lambda_{1,2},\lambda_{1,3};\lambda_{2,1},\lambda_{3,1}) \sim \sin \lambda_{1,2} \cos \lambda_{1,2} \sin \lambda_{1,3} \cos^3 \lambda_{1,3} d \lambda_{1,1}d\lambda_{1,2}d\lambda_{1,3}d\lambda_{2,1}d\lambda_{3,1} \ .
\end{align*}
It now remains to find a general expression for $\left| \det M_1 \right|$ for arbitrary $d$, since the formula of $J_d$ can then be derive via (\ref{recursion}) and induction. In general, the matrix $M_1$ does not have a simple form which makes it difficult to compute its determinant. Here, it is useful to exploit that, first, there is no explicit dependence of $J_d$ on parameters corresponding to phase operations, i.e. all $\lambda_{m,n}$ with $m\geq n$. This can directly be followed from the construction of the composite parameterization in combination with the Haar measure on $\mathcal{U}(2)$ and the Jacobian matrix (\ref{jacobianU2}). Since in the Jacobian matrix (\ref{jacobianU2}) and in its determinant the phase operations $\lambda_{1,1},\lambda_{2,1},\lambda_{2,2}$ only appear as $e^{i\lambda_{1,1}},e^{i\lambda_{2,1}},e^{i\lambda_{2,2}}$, i.e. complex numbers of magnitude 1, and since in the end we are only interested in the absolute value of this determinant, it is clear that $dU_2$ (\ref{U2}) is independent of these parameters. As the composite parameterization only combines $\mathcal{U}(2)$-operations by incorporating all possible $2$-dimensional subspaces in $\mathbb{C}^d$ the same statement holds of course true for all $d$ (see for instance (\ref{U3}) for the case $d=3$). Second, the circumstance that $g(d)$ is independent of all parameters $\lambda_{m,n}$ with $\min\{m,n\}\geq2$ implies that these parameters do not affect $\left| \det M_1 \right|$\footnote{The independence of $\left| \det M_1 \right|$ on $\lambda_{m,n}$ with $\min\{m,n\}\geq2$ can also be confirmed by exploiting again the left and right invariance of the Haar measure.}. For these reasons one can set the parameters $\lambda_{m,n}$ with $\min\{m,n\}\geq2$ and $m\geq n$ to zero in (\ref{drotU}) without altering $|\det M_1|$, i.e.
\begin{align}
\label{Moverline}
|\det M_1|=\left|\det \left(\left. M_1 \right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}}\right)\right| \equiv |\det \overline{M}_1| \ .
\end{align}
Here, (\ref{derivs}) reduces to
\begin{align*}
\left.U_{1,y}^{\dagger}Y_{1,y}U_{1,y}\right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}}=&\left[ \prod_{n=y+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right]^{\dagger} Y_{1,y} \left[ \prod_{n=y+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right] \ , \\
\equiv & O_{1,y}^{T}Y_{1,y} O_{1,y}\\
\left.U_{x,1}^{\dagger}P_{x}U_{x,1}\right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}} =&\left[\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right]^{\dagger} P_{x} \left[\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right] \ ,\\
\equiv & O_{x,1}^{T}P_{x}O_{x,1} \ ,
\end{align*}
where $O_{x,1}$ and $O_{1,y}$ are orthogonal (real) matrices since they are a product of the operations $\exp(i Y_{1,n} \lambda_{1,n})$ which explicitly read
\begin{align}
\label{explicit}
\cos (\lambda_{1,n})\ket{1}\bra{1} + \sin (\lambda_{1,n})\ket{1}\bra{n}-\sin(\lambda_{1,n})\ket{n}\bra{1}+ \cos (\lambda_{m,n})\ket{n}\bra{n} + \sum_{k \neq 1,n}& \ket{k}\bra{k} \ .
\end{align}
According to (\ref{basisorder}), the $2(d-1) \times 2 (d-1)$ elements of $\overline{M}_1$ are now determined by
\begin{align}
\label{Yexp}
\Tr\left(b^{\dagger}_k O_{1,y}^{T}Y_{1,y}O_{1,y}\right)=\Tr\left(O_{1,y}b^{\dagger}_k O_{1,y}^{T}Y_{1,y}\right)
\end{align}
and
\begin{align}
\label{Pexp}
\Tr\left(b^{\dagger}_k O_{x,1}^{T}P_{x}O_{x,1}\right)=\Tr\left(O_{x,1}b^{\dagger}_k O_{x,1}^{T}P_{x}\right)
\end{align}
with $x,y\in \{2,\ldots,d\}$, and
\begin{align}
\label{m1basis}
b_k=b^{\dagger}_k\, = \,\left\{\begin{array}{lll} \ket{1}\bra{(k+3)/2}+\ket{(k+3)/2}\bra{1} &= \ X_{1,(k+3)/2} \hspace{0.3cm} & k \ \mbox{- odd} \ ,\\
-i\ket{1}\bra{(k+2)/2}+i\ket{(k+2)/2}\bra{1} &= \ Y_{1,(k+2)/2} & k \ \mbox{- even} \ , \end{array}\right.
\end{align}
with $k\in\{1,\ldots,2(d-1)\}$. Now consider the coefficients (\ref{Yexp}) and (\ref{Pexp}) depending on different $X_{1,m}$ and $Y_{1,m} \ (2 \leq m \leq d)$ of (\ref{m1basis}):\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{1,y}X_{1,m} O_{1,y}^{T}Y_{1,y}\right)$ for arbitrary $m$:
\begin{align}
\Tr\left(O_{1,y}X_{1,m} O_{1,y}^{T}Y_{1,y}\right)&=0 \ .
\end{align}
\hspace{0.2cm}Note: $O_{1,y}X_{1,m} O_{1,y}^{T}$ is symmetric while $Y_{1,y}$ is antisymmetric.\\
\vspace{-0.3cm}\\$\bullet$ $\Tr\left(O_{1,y}Y_{1,m} O_{1,y}^{T}Y_{1,y}\right)$ for $m < y$:
\begin{align}
\Tr\left(O_{1,y}Y_{1,m} O_{1,y}^{T}Y_{1,y}\right)=0 \ .
\end{align}
\hspace{0.2cm}Note: $O_{1,y}Y_{1,m} O_{1,y}^{T}$ is orthogonal to $Y_{1,y}$ for $m < y$ since the product in\\
$O_{1,y}=\prod_{n=y+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})$ starts with $n=y+1$.\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{x,1}Y_{1,m} O_{x,1}^{T}P_{x}\right)$ for arbitrary $m$:
\begin{align}
\Tr\left(O_{x,1}Y_{1,m} O_{x,1}^{T}P_{x}\right)&=\bra{x}O_{x,1}Y_{1,m} O_{x,1}^{T}\ket{x} \ ,\\
&=-i\bra{x}O_{x,1}\ket{1}\bra{m} O_{x,1}^{T}\ket{x}+i\bra{x}O_{x,1}\ket{m}\bra{1} O_{x,1}^{T}\ket{x} \ ,\\
&=0 \ .
\end{align}
\hspace{0.2cm}Note: $O_{x,1}$ is an orthogonal (real) matrix.\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{x,1}X_{1,m} O_{x,1}^{T}P_{x}\right)$ for $m < x$:
\begin{align}
\Tr\left(O_{x,1}X_{1,m} O_{x,1}^{T}P_{x}\right)&=\bra{x}O_{x,1}X_{1,m} O_{x,1}^{T}\ket{x} \ , \\
&=\bra{x}O_{x,1}\ket{1}\bra{m} O_{x,1}^{T}\ket{x}+\bra{x}O_{x,1}\ket{m}\bra{1} O_{x,1}^{T}\ket{x} \ ,\\
&=0 \ .
\end{align}
\hspace{0.2cm}Note: $O_{x,1}=\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})$ has no off-diagonal elements $\bra{x}O_{x,1}\ket{m}$ for $m<x$.\\
\vspace{-0.3cm}\\
These four observations imply for the operator basis and parameter order
\begin{align}
\begin{array}{cccccccc}
b_1&=&\ket{1}\bra{2}+\ket{2}\bra{1}\ ,& \hspace{1cm} &\alpha_1&=&\lambda_{2,1}\ ,\\
b_2&=&-i\ket{1}\bra{2}+i\ket{2}\bra{1}\ ,& \hspace{1cm} &\alpha_2&=&\lambda_{1,2}\ ,\\
b_3&=&\ket{1}\bra{3}+\ket{3}\bra{1}\ ,& \hspace{1cm} &\alpha_3&=&\lambda_{3,1}\ ,\\
b_4&=&-i\ket{1}\bra{3}+i\ket{3}\bra{1}\ ,& \hspace{1cm} &\alpha_4&=&\lambda_{1,3}\ ,\\
&\vdots& & & &\vdots\\
b_{2(d-1)-1}&=&\ket{1}\bra{d}+\ket{d}\bra{1}\ ,& \hspace{1cm} &\alpha_{2(d-1)-1}&=&\lambda_{d,1}\ ,\\
b_{2(d-1)}&=&-i\ket{1}\bra{d}+i\ket{d}\bra{1}\ ,& \hspace{1cm} &\alpha_{2(d-1)}&=&\lambda_{1,d}\ ,\\
\end{array}
\end{align}
that $\overline{M}_1$ is a lower triangular matrix\\
\setlength{\unitlength}{1cm}
\begin{picture}(14,5)(0,0)
\put(5,2){\fontsize{12}{12}\selectfont\makebox(0,0)[]{$\overline{M}_1= (j'_{k,l})_{\mbox{\tiny{$k,l=1,\ldots,2(d-1)$}}} =$ \strut}}
\put(8.5,0){\includegraphics[scale=0.2]{m1.eps}}
\put(13.5,0.4){\fontsize{12}{12}\selectfont\makebox(0,0)[]{$.$ \strut}}
\end{picture}\\
For the determinant we can now restrict on determining the diagonal entries of $\overline{M}_1$ which are given by\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{1,y}Y_{1,m} O_{1,y}^{T}Y_{1,y}\right)/\Tr(Y^2_{1,m})$ for $m = y$:
\begin{align}
\Tr\left(O_{1,m}Y_{1,m} O_{1,m}^{T}Y_{1,m}\right)=&\bra{1}O_{1,m}Y_{1,m} O_{1,m}^{T}Y_{1,m}\ket{1}+\bra{m}O_{1,m}Y_{1,m} O_{1,m}^{T}Y_{1,m}\ket{m} \ ,\\
=&i\bra{1}O_{1,m}Y_{1,m} O_{1,m}^{T}\ket{m}-i\bra{m}O_{1,m}Y_{1,m} O_{1,m}^{T}\ket{1} \ , \\
=&\bra{1}O_{1,m}\ket{1}\bra{m} O_{1,m}^{T}\ket{m}-\bra{1}O_{1,m}\ket{m}\bra{1} O_{1,m}^{T}\ket{m} \nonumber \\
&-\bra{m}O_{1,m}\ket{1}\bra{m} O_{1,m}^{T}\ket{1}+\bra{m}O_{1,m}\ket{m}\bra{1} O_{1,m}^{T}\ket{1} \ , \\
=&2\bra{1}O_{1,m}\ket{1}\bra{m} O_{1,m}\ket{m} \ . \label{coeff2}
\end{align}
\hspace{0.2cm}Note: $O_{1,m}=\prod_{n=m+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})$ does not have off-diagonal element $\bra{1}O_{1,m}\ket{m}$ since the product starts with $n=m+1$.\\
$\bullet$ $\Tr\left(O_{x,1}X_{1,m} O_{x,1}^{T}P_{x}\right)/\Tr(X^2_{1,m})$ for $m = x$:
\begin{align}
\Tr\left(O_{m,1}X_{1,m} O_{m,1}^{T}P_{m}\right)&=\bra{m}O_{m,1}X_{1,m} O_{m,1}^{T}\ket{m} \ , \\
&=\bra{m}O_{m,1}\ket{1}\bra{m}O_{m,1}^{T}\ket{m}+\bra{m}O_{m,1}\ket{m}\bra{1} O_{m,1}^{T}\ket{m} \ , \\
&=2\bra{m}O_{m,1}\ket{1}\bra{m}O_{m,1}\ket{m} \ . \label{coeff1}
\end{align}
\vspace{-0.3cm}\\
We have thus found simple relations between the diagonal entries of $\overline{M}_1$ and the matrix elements ($2\leq m \leq d$)
\begin{align}
\bra{1}O_{1,m}\ket{1}&=\prod_{n=m+1}^{d}\cos \lambda_{1,n} \ , \\
\bra{m} O_{1,m}\ket{m}&=1 \ , \\
\label{rel1}
\bra{m}O_{m,1}\ket{1}&=-\sin \lambda_{1,m}\prod_{n=m+1}^{d}\cos \lambda_{1,n} \ , \\
\label{rel2}
\bra{m}O_{m,1}\ket{m}&=\cos \lambda_{1,m} \ ,
\end{align}
which can easily be obtained via the definitions of $O_{m,1}$ and $O_{1,m}$ together with (\ref{explicit}). Hence, since $\Tr(X^2_{1,m})=\Tr(Y^2_{1,m})=2$, we find that the diagonal entries of $\overline{M}_1$ are
\begin{align}
\label{diagelements1}
\Tr\left(O_{1,m}Y_{1,m} O_{1,m}^{T}Y_{1,m}\right)/\Tr(Y^2_{1,m})&=\prod_{n=m+1}^{d}\cos \lambda_{1,n} \ , \\
\label{diagelements2}
\Tr\left(O_{m,1}X_{1,m} O_{m,1}^{T}P_{m}\right)/\Tr(X^2_{1,m})&=-\sin \lambda_{1,m}\prod_{n=m}^{d}\cos \lambda_{1,n} \ .
\end{align}
If we multiply all these entries ($m=2,\ldots,d$) we finally obtain
\begin{align}
\left|\det \overline{M}_1 \right|&= \left[\prod_{k=2}^{d}\left(\prod_{n=k+1}^{d}\cos \lambda_{1,n}\right)\right] \cdot \left[\prod_{l=2}^{d}\left( \sin \lambda_{1,l}\prod_{n=l}^{d}\cos \lambda_{1,n}\right) \right] \ , \\
&=\prod_{n=2}^{d}\sin(\lambda_{1,n})\cos^{2(n-1)-1}(\lambda_{1,n}) \ .
\end{align}
Using the recursion formula (\ref{recursion}) and induction one finds the final result
\begin{align}
\label{finalJd}
J_d=\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \ .
\end{align}
The corresponding normalizing constant $N_d$ is given by the integral
\begin{align}
N_d=&\int_{\mathcal{U}(d)}J_d\prod_{k,l=1}^{d}d\lambda_{k,l} \ , \\
=&\int_{\lambda_{k,l}=0}^{2\pi \ (k\geq l)}\int_{\lambda_{k,l}=0}^{\pi/2 \ (k<l)} \left[\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \right] \prod_{k,l=1}^{d}d\lambda_{k,l} \ ,
\end{align}
which can straightforwardly be computed
\begin{align}
N_d=&(2\pi)^{d(d+1)/2} \prod_{m=1}^{d-1}\prod_{n=m+1}^{d} \left. \frac{-1}{2(n-m)}\cos^{2(n-m)}(\lambda_{m,n}) \right|_{\lambda_{m,n}=0}^{\pi/2} \ , \\
=&\frac{(2\pi)^{d(d+1)/2}}{\prod_{m=1}^{d-1}\prod_{n=m+1}^{d} 2(n-m)} \ .
\end{align}
\end{proof}
\section{Haar measure on the special unitary group $\mathcal{SU}(d)$}
\label{haarSU}
\begin{samepage}
\begin{thm}
\label{theorem4}
In terms of the $d^2-1$ parameters $\lambda_{m,n}$ introduced in Theorem $2$ the (normalized) Haar measure on the special unitary group $\mathcal{SU}(d)$ reads
\begin{align}
\label{HaarSU}
dU_{d}=\frac{1}{N_d}\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \prod_{k,l}d\lambda_{k,l} \ ,
\end{align}
with
\begin{align}
N_d=\frac{2^{d-1}\pi^{d(d+1)/2-1}}{\prod_{m=1}^{d-1}\prod_{n=m+1}^{d} 2(n-m)}
\end{align}
such that $\int_{\mathcal{SU}(d)} dU_{d} =1$.
\end{thm}
\end{samepage}
\begin{proof}
The proof is similar to the previous one --- only minor modifications have to be made. Given the special unitary group $\mathcal{SU}(d)$ in parameterized form $U(\alpha_1,\ldots,\alpha_{d^2-1})$, to construct the Haar measure one must determine the absolute value $J_d = \left| \det (j_{k,l}) \right|$ of the determinant of the Jacobian matrix
\begin{align}
(j_{k,l})=\frac{\partial(u_1,\ldots,u_{d^2-1})}{\partial (\alpha_1,\ldots,\alpha_{d^2-1})} \ ,
\end{align}
wherein $\{ u_k \}$ are coefficients of the special unitary operator $U(\alpha_1,\ldots,\alpha_{d^2-1})$ expanded in an orthogonal basis $\{ b_k \}$ of traceless operators\footnote{A basis for the vector space of operators $\mathbb{C}^d \times \mathbb{C}^d$ has $d^2$ elements. The constraint $\det U=1$ on special unitary operators implies that the trace of any derivative $\partial U / \partial \alpha_l$ with respect to an arbitrary parameter is always zero. Traceless operators form a $d^2-1$ dimensional subspace of $\mathbb{C}^d \times \mathbb{C}^d$. As we have $d^2-1$ parameters $\{ \alpha_l \}$ it is required to express the derivatives
$\partial U / \partial \alpha_l$ in a basis of this subspace to obtain $d^2-1$ linearly independent column vectors.} of $\mathbb{C}^d \times \mathbb{C}^d$, i.e.
\begin{align}
u_k=\frac{\Tr(b_k^{\dagger}U)}{\Tr(b_k^{\dagger}b_k)} \hspace{1.8cm} \left( \Rightarrow \frac{\partial U}{\partial \alpha_l}=\sum_{k=1}^{d^2-1} \frac{\partial u_k}{\partial \alpha_l} b_k \right) \ .
\end{align}
In this terminology, the (normalized) Haar measure reads
\begin{align}
dU_d=\frac{J_d}{N_d} \prod_{l=1}^{d^2-1} d\alpha_l \ .
\end{align}
Our aim is to derive a general expression of $dU_d$ for arbitrary $d$ in terms of the parameterization introduced in Section \ref{compSU}, i.e.
\begin{align}
dU_d=\frac{J_d}{N_d} \prod_{k,l}d\lambda_{k,l} \ .
\end{align}
As in the previous proof we make use of the left invariance of the Haar measure on $\mathcal{SU}(d)$, i.e. $J_d=|\det (j_{k,l})|=|\det (j'_{k,l})|$ where
\begin{align}
j_{k,l}=\frac{\Tr(b_k^{\dagger} \frac{\partial U}{\partial \alpha_l})}{\Tr(b_k^{\dagger}b_k)} \hspace{0.5cm} \longrightarrow \hspace{0.5cm} j'_{k,l}=\frac{\Tr(b_k^{\dagger} U_1 \frac{\partial U}{\partial \alpha_l})}{\Tr(b_k^{\dagger}b_k)} \ .
\end{align}
If for the composite parameterization (Theorem \ref{theorem2}) $U_1$ is chosen to be $-iU^{\dagger}_C$ one obtains analogously to (\ref{diagglobal}) -- (\ref{drotU}) that
\begin{align}
\label{derivs2}
-iU_C^{\dagger}\frac{\partial U_C}{\partial \lambda_{x,y}}\, = \,\left\{\begin{array}{ll} U_{x,y}^{\dagger}Y_{x,y}U_{x,y} \hspace{1cm} & \mbox{for} \ \ x<y \\
\hspace{0.7cm} Z_{x,d} & \mbox{for} \ \ x=y \ \\
U_{x,y}^{\dagger}Z_{y,x}U_{x,y} & \mbox{for} \ \ x>y \ , \end{array}\right.
\end{align}
where
\begin{align}
U_{x,y}\, = \,\left\{\begin{array}{ll} \left[\prod_{n=y+1}^{d} \Lambda_{x,n} \right]\left[\prod_{m=x+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=x}^{d-1} \exp(i Z_{l,d} \lambda_{l,l}) & \mbox{for} \ \ x<y \\
\left[\prod_{n=x}^{d} \Lambda_{y,n} \right]\left[\prod_{m=y+1}^{d-1} \prod_{n=m+1}^{d} \Lambda_{m,n} \right] \prod_{l=y}^{d-1} \exp(i Z_{l,d} \lambda_{l,l}) & \mbox{for} \ \ x>y \ . \end{array}\right.
\end{align}
These operators are now to be expressed in an orthogonal basis $\{ b_k \}$ of traceless operators. Since the first $d^2-d$ operators that were used in (\ref{basisorder}) already are mutually orthogonal and traceless, only the $d$ diagonal operators $\ket{k}\bra{k}$ with $1\leq k \leq d$ have to be replaced by $d-1$ diagonal operators with vanishing trace. A convenient choice are the $d-1$ mutually orthogonal operators\footnote{Note that these are the generalized diagonal Gell-Mann matrices \cite{Krammer} in reversed order.}
\begin{align}
L_k=\sqrt{\frac{2}{(d-k)(d-k+1)}} \left( -(d-k)\ket{k}\bra{k} + \sum_{n=k+1}^{d} \ket{n} \bra{n} \right) \ ,
\end{align}
with $1\leq k \leq d-1$ obeying $\Tr(L_k^{\dagger} L_k)=\Tr(L_k^2)=2$. For the order
\begin{align}
\label{basisorderSU}
\begin{array}{cccccccc}
b_1&=&\ket{1}\bra{2}+\ket{2}\bra{1}\ ,& &\alpha_1&=&\lambda_{2,1}\ ,\\
b_2&=&-i\ket{1}\bra{2}+i\ket{2}\bra{1}\ ,& &\alpha_2&=&\lambda_{1,2}\ ,\\
b_3&=&\ket{1}\bra{3}+\ket{3}\bra{1}\ ,& &\alpha_3&=&\lambda_{3,1}\ ,\\
b_4&=&-i\ket{1}\bra{3}+i\ket{3}\bra{1}\ ,& &\alpha_4&=&\lambda_{1,3}\ ,\\
&\vdots& & & &\vdots\\
b_{2(d-1)-1}&=&\ket{1}\bra{d}+\ket{d}\bra{1}\ ,& &\alpha_{2(d-1)-1}&=&\lambda_{d,1}\ ,\\
b_{2(d-1)}&=&-i\ket{1}\bra{d}+i\ket{d}\bra{1}\ ,& &\alpha_{2(d-1)}&=&\lambda_{1,d}\ ,\\
b_{2(d-1)+1}&=&\ket{2}\bra{3}+\ket{3}\bra{2}\ ,& &\alpha_{2(d-1)+1}&=&\lambda_{3,2}\ ,\\
b_{2(d-1)+2}&=&-i\ket{2}\bra{3}+i\ket{3}\bra{2}\ ,& &\alpha_{2(d-1)+2}&=&\lambda_{2,3}\ ,\\
b_{2(d-1)+3}&=&\ket{2}\bra{4}+\ket{4}\bra{2}\ ,& &\alpha_{2(d-1)+3}&=&\lambda_{4,2}\ ,\\
b_{2(d-1)+4}&=&-i\ket{2}\bra{4}+i\ket{4}\bra{2}\ ,& &\alpha_{2(d-1)+4}&=&\lambda_{2,4}\ ,\\
&\vdots& & & &\vdots\\
b_{d^2-d-1}&=&\ket{d-1}\bra{d}+\ket{d}\bra{d-1}\ ,& &\alpha_{d^2-d-1}&=&\lambda_{d,d-1}\ ,\\
b_{d^2-d}&=&-i\ket{d-1}\bra{d}+i\ket{d}\bra{d-1}\ ,& &\alpha_{d^2-d}&=&\lambda_{d-1,d}\ ,\\
b_{d^2-d+1}&=&L_1\ ,& \hspace{1cm} &\alpha_{d^2-d+1}&=&\lambda_{1,1}\ ,\\
b_{d^2-d+2}&=&L_2\ ,& &\alpha_{d^2-d+2}&=&\lambda_{2,2}\ ,\\
&\vdots& & & &\vdots\\
b_{d^2-1}&=& L_{d-1}\ ,& &\alpha_{d^2-1}&=&\lambda_{d-1,d-1}\ ,& \\
\end{array}
\end{align}
the $(d^2-1) \times (d^2-1)$ Jacobian matrix is again lower block-triangular\\
\setlength{\unitlength}{1cm}
\begin{picture}(14,9)(0,0)
\put(3,4.5){\fontsize{12}{12}\selectfont\makebox(0,0)[]{$(j'_{k,l}) \ = $ \strut}}
\put(5,0){\includegraphics[scale=0.4]{jacobian2.eps}}
\put(14,0){\fontsize{12}{12}\selectfont\makebox(0,0)[]{$.$ \strut}}
\end{picture}\\
Here, the block $D$ is not diagonal but lower triangular since
\begin{align}
\Tr(L_k^{\dagger}Z_{m,d})=\,\left\{\begin{array}{ll} 0 \hspace{1cm} & \mbox{for} \ \ k<m \\
\sqrt{\frac{2}{(d-k)(d-k+1)}} (k-d-1) & \mbox{for} \ \ k=m \ \\
-\sqrt{\frac{2}{(d-k)(d-k+1)}} & \mbox{for} \ \ k>m \ . \end{array}\right.
\end{align}
Since the diagonal entries of $D$ are merely real numbers we find again that
\begin{align}
J_d=\left| \det (j'_{k,l}) \right|= c_d \prod_{i=1}^{d-1}\left| \det M_i \right| \ ,
\end{align}
where $c_d$ is a constant that can be dropped since the Haar measure will be normalized at the end anyhow. As the composite parameterization of the unitary group and the special unitary group share the same structure the discussion between (\ref{recursion}) -- (\ref{Moverline}) essentially remains the same, meaning that (ignoring irrelevant constants) one finds again the recursion formula
\begin{align}
J_d=\left| \det \overline{M}_1 \right|J_{d-1} \ ,
\end{align}
where $ \overline{M}_1=\left. M_1 \right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}}$ and $\lambda_{m,n} \rightarrow \lambda_{m+1,n+1}$ in $J_{d-1}$. Here, the relevant operators of (\ref{derivs2}) have the form
\begin{align*}
\left.U_{1,y}^{\dagger}Y_{1,y}U_{1,y}\right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}}&=\left[ \prod_{n=y+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right]^{\dagger} Y_{1,y} \left[ \prod_{n=y+1}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right] \ , \\
&\equiv O_{1,y}^{T}Y_{1,y} O_{1,y} \ , \\
\left.U_{x,1}^{\dagger}Z_{1,x}U_{x,1}\right|_{\{\lambda_{m,n}=0|(m,n) \notin \{(1,2),\ldots,(1,d) \} \}}&=\left[\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right]^{\dagger} Z_{1,x} \left[\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})\right] \ , \\
&\equiv O_{x,1}^{T}Z_{1,x}O_{x,1} \ ,
\end{align*}
containing the orthogonal matrices $O_{x,1}$ and $O_{1,y}$ that also appeared in the previous proof. The $2(d-1) \times 2 (d-1)$ elements of $\overline{M}_1$ are determined by (compare with the order (\ref{basisorderSU}))
\begin{align}
\label{Yexp2}
\Tr\left(b^{\dagger}_k O_{1,y}^{T}Y_{1,y}O_{1,y}\right)=\Tr\left(O_{1,y}b^{\dagger}_k O_{1,y}^{T}Y_{1,y}\right)
\end{align}
and
\begin{align}
\label{Zexp2}
\Tr\left(b^{\dagger}_k O_{x,1}^{T}Z_{1,x}O_{x,1}\right)=\Tr\left(O_{x,1}b^{\dagger}_k O_{x,1}^{T}Z_{1,x}\right)
\end{align}
with $x,y\in \{2,\ldots,d\}$, and
\begin{align}
\label{m1blocko}
b_k=b^{\dagger}_k\, = \,\left\{\begin{array}{lll} \ket{1}\bra{(k+3)/2}+\ket{(k+3)/2}\bra{1} &= \ X_{1,(k+3)/2} \hspace{0.3cm} & k \ \mbox{- odd} \\
-i\ket{1}\bra{(k+2)/2}+i\ket{(k+2)/2}\bra{1} &= \ Y_{1,(k+2)/2} & k \ \mbox{- even} \ , \end{array}\right.
\end{align}
with $k\in\{1,\ldots,2(d-1)\}$. Now consider the coefficients of (\ref{Yexp2}) and (\ref{Zexp2}) for the different operator basis elements occurring in (\ref{m1blocko}): First, notice that $\overline{M}_1$ is again lower triangular as in the previous proof since (\ref{Yexp2}) and (\ref{Yexp}) are identical and\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{x,1}Y_{1,m} O_{x,1}^{T}Z_{1,x}\right)=0$ for arbitrary $m$.\\
\vspace{-0.3cm}\\
\hspace{0.2cm}Note: $O_{1,y}Y_{1,m} O_{1,y}^{T}$ is antisymmetric, while $Z_{1,x}$ is symmetric.\\
\vspace{-0.3cm}\\
$\bullet$ $\Tr\left(O_{x,1}X_{1,m} O_{x,1}^{T}Z_{1,x}\right)=0$ for $m < x$:
\begin{align}
\Tr\left(O_{x,1}X_{1,m} O_{x,1}^{T}Z_{1,x}\right)=&\bra{1}O_{x,1}X_{1,m} O_{x,1}^{T}\ket{1}-\bra{x}O_{x,1}X_{1,m} O_{x,1}^{T}\ket{x} \ , \\
=&\bra{1}O_{x,1}\ket{1}\bra{m} O_{x,1}^{T}\ket{1}-\bra{1}O_{x,1}\ket{m}\bra{1} O_{x,1}^{T}\ket{1} \nonumber\\
&-\bra{x}O_{x,1}\ket{1}\bra{m} O_{x,1}^{T}\ket{x}-\bra{x}O_{x,1}\ket{m}\bra{1} O_{x,1}^{T}\ket{x} \ , \\
=&0 \ .
\end{align}
\hspace{0.2cm}Note: $O_{x,1}=\prod_{n=x}^{d} \exp(i Y_{1,n} \lambda_{1,n})$ does not have off-diagonal elements $\bra{1}O_{x,1}\ket{m}$ and $\bra{x}O_{x,1}\ket{m}$ for $m<x$.\\
\vspace{-0.3cm}\\
Thus, it again suffices to compute the diagonal entries of $\overline{M}_1$, half of which are already known from (\ref{diagelements1})
\begin{align}
\label{diag1}
\Tr\left(O_{1,m}Y_{1,m} O_{1,m}^{T}Y_{1,m}\right)/\Tr(Y^2_{1,m})&=\prod_{n=m+1}^{d}\cos \lambda_{1,n} \ .
\end{align}
The remaining ones are found to be
\begin{align}
\label{diag2}
\Tr\left(O_{1,m}X_{1,m} O_{1,m}^{T}Z_{1,m}\right)/\Tr(X^2_{1,m})&=2\sin \lambda_{1,m}\prod_{n=m}^{d}\cos \lambda_{1,n} \ ,
\end{align}
which directly follow from
\begin{align}
\Tr\left(O_{1,m}X_{1,m} O_{1,m}^{T}Z_{1,m}\right)=&\bra{1}O_{m,1}X_{1,m} O_{m,1}^{T}\ket{1}-\bra{m}O_{m,1}X_{1,m} O_{m,1}^{T}\ket{m} \ , \\
=&\bra{1}O_{m,1}\ket{1}\bra{m}O_{m,1}^{T}\ket{1}+\bra{1}O_{m,1}\ket{m}\bra{1} O_{m,1}^{T}\ket{1} \nonumber\\
&-\bra{m}O_{m,1}\ket{1}\bra{m}O_{m,1}^{T}\ket{m}-\bra{m}O_{m,1}\ket{m}\bra{1} O_{m,1}^{T}\ket{m} \ , \nonumber\\
=&2\bra{1}O_{m,1}\ket{m}\bra{1}O_{m,1}\ket{1}-2\bra{m}O_{m,1}\ket{1}\bra{m}O_{m,1}\ket{m} \label{coeffsu1} \ .
\end{align}
and the definition of $O_{m,1}$ together with (\ref{explicit})
\begin{align}
\bra{1}O_{m,1}\ket{m}&=\sin \lambda_{1,m} \ , \\
\bra{1}O_{m,1}\ket{1}&=\prod_{n=m}^{d}\cos \lambda_{1,n} \ , \\
\bra{m}O_{m,1}\ket{1}&=-\sin \lambda_{1,m}\prod_{n=m+1}^{d}\cos \lambda_{1,n} \ , \\
\bra{m}O_{m,1}\ket{m}&=\cos \lambda_{1,m} \ .
\end{align}
Since (\ref{diag2}) differs only by the factor two from (\ref{diagelements1}) we again obtain for $J_d$ the result (\ref{finalJd})
\begin{align}
J_d=c'_d \prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \ ,
\end{align}
up to an irrelevant multiplicative constant $c'_d$. Hence, the Haar measure on $\mathcal{SU}(d)$ reads
\begin{align}
dU_{d}=\frac{1}{N_d}\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \prod_{k,l}d\lambda_{k,l} \ .
\end{align}
The normalizing constant $N_d$ is given by the integral
\begin{align}
N_d=&\int_{\lambda_{k,l}=0}^{2\pi \ (k=l)}\int_{\lambda_{k,l}=0}^{\pi \ (k> l)}\int_{\lambda_{k,l}=0}^{\pi/2 \ (k<l)} \left[\prod_{m=1}^{d-1}\prod_{n=m+1}^{d}\sin(\lambda_{m,n})\cos^{2(n-m)-1}(\lambda_{m,n}) \right] \prod_{k,l}d\lambda_{k,l} \ . \nonumber
\end{align}
Since there are $d-1$ parameters $\lambda_{k,l}$ with $k=l$ and $d(d-1)/2$ parameters $\lambda_{k,l}$ with $k>l$ one finally obtains
\begin{align}
N_d=&2^{d-1}\pi^{d(d+1)/2-1} \prod_{m=1}^{d-1}\prod_{n=m+1}^{d} \left. \frac{-1}{2(n-m)}\cos^{2(n-m)}(\lambda_{m,n}) \right|_{\lambda_{m,n}=0}^{\pi/2} \ , \\
=&\frac{2^{d-1}\pi^{d(d+1)/2-1}}{\prod_{m=1}^{d-1}\prod_{n=m+1}^{d} 2(n-m)} \ .
\end{align}
\end{proof}
\section{Remarks on integrals over unitary groups}
\label{NotesInt}
As previously mentioned, our results can be used to compute group integrals, i.e. integrals of the form $\int f(U,U^*) dU$ where one integrates over the entire group $\mathcal{U}(d)$ or $\mathcal{SU}(d)$, respectively. At least three things are needed when one intends to \emph{explicitly} compute such integrals: A parameterization of the corresponding group, exact knowledge of the parameter ranges and the normalized Haar measure. All this is provided in the present paper. Theorems \ref{theorem1} -- \ref{theorem4} can straightforwardly be applied without knowledge of further technicalities (e.g. details appearing in the proofs). Whether or not a given integral can be solved analytically in this way of course depends on the integrand. However, for many physical problems the function $f(U,U^*)$ is a polynomial in the components of $U$ and $U^{*}$. In this case, when $U$ and $dU$ are inserted in parameterized form according to Theorems \ref{theorem1} -- \ref{theorem4}, we have that the integrand is a polynomial in $\cos \lambda_{m,n}$, $\sin \lambda_{m,n}$ and $e^{\pm i\lambda_{m,n}}$. Neglecting the computational effort, such integrals can always be solved analytically (see Ref.~\cite{Gradst} and references therein). Besides that, our results constitute a good starting point for the integration of non-polynomial functions using numerical methods.
A detailed analysis on integrals that can be solved in this way shall be presented in a subsequent paper. We are convinced that due to the simplicity of the parameterization and the associated Haar measure, it is possible to find several general results and to gain a better understanding of integrals over $\mathcal{U}(d) \ / \ \mathcal{SU}(d)$. In order not to go beyond the scope of this paper we conclude with simple examples that can be compared with existing results as a consistency check.
\subsection{Example 1}
In Ref.~\cite{Collins} it was shown that
\begin{align}
\label{collinsrel}
\int_{\mathcal{U}(d)} |\bra{1}U\ket{1}|^4 dU=\frac{2}{d(d+1)} \ .
\end{align}
By means of our results, it is straightforward to analytically confirm this relation and to find the general solution for $\int |\bra{1}U\ket{1}|^p dU$ for arbitrary $p \in \mathbb{N}$. In parameterized form we have $\bra{1}U_C\ket{1}=e^{i \lambda_{1,1}} \prod_{n=2}^{d} \cos(\lambda_{1,n})$, and accordingly $|\bra{1}U_C\ket{1}|^p=\prod_{n=2}^{d} \cos^p(\lambda_{1,n})$ only depends on $\lambda_{1,2},\ldots,\lambda_{1,d}$. Due to this, the integral simplifies as follows (see also Appendix \ref{diffmatrix})
\begin{align}
\label{bintegral}
&\int_{\mathcal{U}(d)} |\bra{1}U\ket{1}|^p dU \ , \\
=&\int_{\lambda_{k,l}=0}^{2\pi \ (k\geq l)}\int_{\lambda_{k,l}=0}^{\pi/2 \ (k<l)} \prod_{n=2}^{d} \cos^p(\lambda_{1,n}) \ dU_d \ , \\
=&\int_{\lambda_{1,2}=0}^{\pi/2}\cdots\int_{\lambda_{1,d}=0}^{\pi/2} \prod_{n=2}^{d} \cos^p(\lambda_{1,n}) 2(n-1) \sin(\lambda_{1,n})\cos^{2(n-1)-1}(\lambda_{1,n}) d\lambda_{1,2} \cdots d\lambda_{1,d} \ , \nonumber \\
=&\int_{\lambda_{1,2}=0}^{\pi/2}\cdots\int_{\lambda_{1,d}=0}^{\pi/2} \prod_{n=2}^{d} 2(n-1) \sin(\lambda_{1,n})\cos^{2(n-1)-1+p}(\lambda_{1,n}) d\lambda_{1,2} \cdots d\lambda_{1,d} \ .
\end{align}
This integral can easily be solved, i.e.
\begin{align}
=&\left. \left[\prod_{n=2}^{d} 2(n-1) \frac{-\cos^{2(n-1)+p}(\lambda_{1,n})}{2(n-1)+p}\right] \right|_{\lambda_{1,n}=0}^{\pi/2} \ , \\
=&\prod_{n=2}^{d} \frac{2(n-1)}{2(n-1)+p} \ .
\label{eintegral}
\end{align}
For the special case $p=4$ the general solution
\begin{align}
\int_{\mathcal{U}(d)} |\bra{1}U\ket{1}|^p dU=\prod_{n=2}^{d} \frac{2(n-1)}{2(n-1)+p}
\end{align}
simplifies to $\prod_{n=2}^{d} \frac{2(n-1)}{2(n+1)}=\frac{2}{d(d+1)}$ as in agreement with Ref.~\cite{Collins}. Note that in this way we have found a simple necessary criterion for testing if a set of matrices constitutes a unitary design. Namely, as there are no distinguished matrix elements and since $|\bra{1}U\ket{1}|^{2t}=\bra{1}U\ket{1}^{t}\bra{1}U^*\ket{1}^{t}$ it holds: A set of unitaries $\{ U_i \}_N$ is a unitary $t$-design only if
\begin{align}
\sum_{i=1}^{N} w_i |\bra{k} U_i \ket{l}|^{2t} =\prod_{n=2}^{d} \frac{n-1}{n-1+t}
\end{align}
for all $k, l \in \{1,\ldots,d\}$, where $w_i$ is the weighting of $U_i$ ($w_i=\frac{1}{N}$ for unweighted designs). Further criteria can be constructed analogously.
\subsection{Example 2}
It is known that bilateral twirling \mbox{$\int U \otimes U \rho U^{\dagger} \otimes U^{\dagger} dU$} of a state $\rho$ on $\mathbb{C}^d \otimes \mathbb{C}^d$ results in a \emph{Werner state} \cite{Werner,Eggeling,Chruscinski,Clarisse} which has the form
\begin{align}
\rho_W=\frac{\mathbbm{1} + \beta (\sum_{i,j=1}^{d} \ket{ij}\bra{ji})}{d(d+\beta)} \ ,
\end{align}
where $-1 \leq \beta \leq 1$. Using a symbolic computation software, we explicitly and analytically twirled the maximally entangled state $\ket{\Psi}=\frac{1}{\sqrt{d}} \sum_{i=1}^{d} \ket{i}\otimes \ket{i}$ of dimensions $d=2,3,4,5$ utilizing Theorem \ref{theorem1} and \ref{theorem3}
\begin{align}
\int_{\lambda_{k,l}=0}^{2\pi \ (k\geq l)}\int_{\lambda_{k,l}=0}^{\pi/2 \ (k<l)} U_C \otimes U_C \ket{\Psi}\bra{\Psi} U_C^{\dagger} \otimes U_C^{\dagger} dU_d \ ,
\end{align}
and alternatively with Theorem \ref{theorem2} and \ref{theorem4}
\begin{align}
\int_{\lambda_{k,l}=0}^{2\pi \ (k=l)}\int_{\lambda_{k,l}=0}^{\pi \ (k> l)}\int_{\lambda_{k,l}=0}^{\pi/2 \ (k<l)} U_C \otimes U_C \ket{\Psi}\bra{\Psi} U_C^{\dagger} \otimes U_C^{\dagger} dU_d \ .
\end{align}
In all cases, this yielded a \emph{Werner state} with $\beta=1$, demonstrating once more the operationality and validity of our results. Note that our results also enable analytical twirling of multipartite qudit states.
\subsection{Example 3}
The entanglement of a bipartite qudit state $\ket{\psi} \in \mathcal{H}_A \otimes \mathcal{H}_B = \mathbb{C}^d \otimes \mathbb{C}^d$ can be quantified via the (normalized) concurrence \cite{conc1,conc2,conc3,conc4} determined by $C^2(\ket{\psi})=\frac{d}{d-1}(1-\Tr(\rho_B^2))$, where $\rho_B$ is the reduced density matrix $\rho_B=\Tr_A(\ket{\psi}\bra{\psi})$. An interesting property of a quantum system is the \emph{a priori entanglement}, i.e. a characteristic such as the generic probability that a state is entangled or the average amount of entanglement over all states in dependence of the dimension $d$. Here, we focus on the average
\begin{align}
\langle C^2 \rangle=\int_{\mathcal{U}(d^2)}C^2(U \ket{\psi}) dU \ ,
\end{align}
where $\ket{\psi}$ is an arbitrary state of $\mathbb{C}^d \otimes \mathbb{C}^d$. If we express $\ket{\psi'}\bra{\psi'}=U \ket{\psi}\bra{\psi} U^{\dagger}$ as $\ket{\psi'}\bra{\psi'}=\sum_{i,j=1}^{d^2} u_{i}u^{*}_{j} \ket{i}\bra{j}$ we obtain the reduced density matrix
\begin{align}
\label{reddensity}
\rho_B=\sum_{i,j=1}^{d} \left(\sum_{n=0}^{d-1} u_{i+nd}u^{*}_{j+nd} \right)\ket{i}\bra{j} \ .
\end{align}
To solve the integral $\int \frac{d}{d-1}(1-\Tr(\rho_B^2))dU$ we can associate each $u_{i}$ with a matrix element of a $d^2 \times d^2$ unitary matrix. In this way, the average $\langle C^2 \rangle$ reduces to a sum of integrals over the polynomials $I_4=\int |\bra{k}U\ket{l}|^4 dU$ and $I_{2,2}=\int |\bra{k}U\ket{l}|^2|\bra{m}U\ket{n}|^2 dU$ where $\bra{k}U\ket{l}$ and $\bra{m}U\ket{n}$ denote distinct matrix elements. Now, taking account of (\ref{reddensity}), it is a simple combinatorial problem to show that in $\int \Tr(\rho_B^2)dU$ the term $I_4$ appears $d^2$ times and $I_{2,2}$ appears $2(d-1)d^2$ times. Hence,
\begin{align}
\langle C^2 \rangle=\frac{d}{d-1}(1-d^2I_4-2(d-1)d^2I_{2,2}) \ .
\end{align}
From (\ref{collinsrel}) we already know that $I_4=\frac{2}{d^2(d^2+1)}$ for integrals over $\mathcal{U}(d^2)$. It remains to determine $I_{2,2}$ for two arbitrary but distinct matrix elements $\bra{k}U\ket{l}$ and $\bra{m}U\ket{n}$. By computing
\begin{align}
I_{2,2}=\int_{\mathcal{U}(d^2)} |\bra{1}U_C\ket{1}|^2|\bra{d^2}U_C\ket{1}|^2 dU_{d^2} \ ,
\end{align}
analogously to (\ref{bintegral}) -- (\ref{eintegral}) one finds with $|\bra{d^2} U_C \ket{1}|^2=\sin^2(\lambda_{1,d^2})$ that $I_{2,2}=\frac{1}{d^2(d^2+1)}=\frac{1}{2}I_4$. Consequently, the average entanglement of a bipartite qudit system is
\begin{align}
\langle C^2 \rangle=\frac{d(d-1)}{d^2+1} \ .
\label{apriorient}
\end{align}
This result is graphically depicted in Figure~\ref{averagecon}. Physically interpreted it means that the higher-dimensional the system is, the more likely it becomes to obtain a highly entangled state when picking a pure state of $\mathcal{H}=\mathbb{C}^d \otimes \mathbb{C}^d$ at random.
\begin{figure}[h]
\centering
\setlength{\unitlength}{1cm}
\begin{picture}(10,6.5)(0,0)
\put(0,0){\includegraphics[scale=1]{concurrence.eps}}
\put(0.5,6.3){\fontsize{12}{12}\selectfont \makebox(0,0)[]{$\langle C^2 \rangle$\strut}}
\put(9.2,0.5){\fontsize{12}{12}\selectfont \makebox(0,0)[]{$d$\strut}}
\end{picture}
\caption{The average entanglement (\ref{apriorient}) of a bipartite qudit system measured by the (squared) concurrence $C^2(\ket{\psi})$ for the dimensions $d=2,\ldots,12$. The average $\langle C^2 \rangle$ increases with the size of the Hilbert space $\mathcal{H}_A \otimes \mathcal{H}_B=\mathbb{C}^d \otimes \mathbb{C}^d$.}
\label{averagecon}
\end{figure}
\section{Summary}
In this paper we adopted the concept of the composite parameterization of the unitary group $\mathcal{U}(d)$ to the special unitary group $\mathcal{SU}(d)$. We showed that both parameterizations can be used equivalently to describe orthonormal vectors and subspaces with the minimal number of parameters. The introduced parameterizations are completely factorized and therefore beneficial for numerical optimizations. We also determined the infinitesimal volume element in terms of the introduced parameters. We derived a general formula of the normalized Haar measure for both the unitary $\mathcal{U}(d)$ and the special unitary group $\mathcal{SU}(d)$ of arbitrary dimension. The found expressions give theoretical insights into the differential structure of $\mathcal{U}(d)$ and $\mathcal{SU}(d)$. Moreover, the Haar measure plays an important role in all kinds of unbiased randomizations. It was stressed that our results also constitute a framework for computing high-order group integrals. By means of our approach, we analytically solved several exemplary integrals and found that the solutions are in agreement with the literature. As integrals over unitary groups appear in various fields from particle physics to quantum optics to quantum information, it is to be expected that our results will find several interesting applications.
\vspace{0.5cm}\\
\textbf{Acknowledgments:}
The authors want to thank Jan Bouda, Artem Kaznatcheev, Patrick Ludl, David Rottensteiner and the Anonymous Referee for helpful remarks and their valuable comments on the manuscript. Christoph Spengler and Marcus Huber acknowledge financial support from the Austrian Science Fund (FWF) - Project P21947N16.
|
\section{Introduction}
The standard $\Lambda$CDM cosmological model suggests a bottom-up sequence of structure formation,
in which a series of merging events culminates in massive clusters of galaxies, the latest structures to form in the observable Universe \citep{Bond1991,Lacey1993}.
The number of clusters as a function of their mass (the steep, high end of the mass function) depends sensitively upon cosmological parameters \citep[e.g.][]{Vikhlinin2009}
and has become an important observational test of cosmology.
Most measurements of cluster masses rely upon the calibration of more easily observable proxies, such as X-ray luminosity, temperature or galaxy richness.
However, clusters form through multiple, dynamic accretions, so are likely to be turbulent places, and the turmoil affects those observable proxies.
It is therefore vital to quantitatively understand the merging process, for example by mapping the distribution of dark matter, stars and baryonic gas in
systems at many different stages of the merger.
Merging clusters of galaxies have become useful laboratories in which to study the nature and interaction properties of dark matter.
The best-studied example is
the \textit{Bullet Cluster} 1ES~0657-558 ($z=0.296$) \citep{Tucker1998,Markevitch2002}. Combined X-ray and gravitational lensing analyses
show a clear separation between the centres of X-ray emission and the peaks in surface mass density -- indicating a fundamental difference between
baryonic gas, which feels the pressure of the collision, and dark matter, which is nearly collisionless \citep{Clowe2004,Clowe2006,Bradav2006}.
The discovery and interpretation of the \textit{Bullet Cluster} has inspired a lively debate about whether
such a system could exist in different cosmological models \citep[see e.g.][]{Hayashi2006}.
Improvements are continuing in numerical simulations \citep{Milosavljevi'c2007,Springel2007,Mastropietro2008,Lee2010}.
Observations have also broadened, with discoveries of a possible line-of-sight merger CL0024+1654
\citep{Czoske2002,Hoekstra2007,Jee2007,Zitrin2009,Umetsu2010,Zu2009,Zu2009a},
and other systems including the \textit{Baby Bullet} (MACS~J0025.4-1222) \citep{Bradav2008},
the \textit{Cosmic Train Wreck} (Abell~520) \citep{Mahdavi2007,Okabe2008}, Abell~2146 \citep{Russell2010},
Abell~521 \citep{Giacintucci2008,Okabe2010a} and Abell~3667 \citep{Finoguenov2010}.
All these systems place potentially tight constraints on the interaction between baryons and dark matter,
and are exemplary probes for our understanding of structure formation within gravitationally bound systems.
In the archival data of 38 merging clusters, \citet{Shan2010} found the
largest offset between X-ray and lensing signals to occur in the massive ($L_X=3.1\times$10$^{45}$ erg/s in the 2-10 keV range, \citet{Allen1998})
cluster Abell~2744 (also known as AC118, or RXCJ0014.3-3022)
at a redshift of $z=0.308$ \citep{Couch1984}.
The $\sim 250$ kpc offset is larger than that in the \textit{Bullet Cluster} -- although, as we shall discuss later,
this value does not describe the separation of the main mass-clump from its stripped gas component.
Nevertheless, Abell~2744 is undergoing a particularly interesting merger.
The complex interplay between multiple dark matter and baryonic components appears to have unleashed `ghost', `dark', `stripped' and `bullet' clusters.
The first hint that Abell~2744 is in the middle of a major merging event arose from observation of a powerful and extended radio halo
($P(1.4 \textrm{GHz}) > 1.6\times 10^{36}$ Watt,~
\citet{Giovannini1999,Govoni2001,Govoni2001a}).
This indicated the presence of relativistic electrons accelerated through
high Mach shocks or turbulence \citep[e.g.][]{Sarazin2004}.
The picture was clarified by X-ray studies \citep{Kempner2004,Zhang2004} that
revealed substructure near the cluster core, plus an additional luminous structure towards the Northwest.
Kinematic observations of cluster member galaxies \citep{Girardi2001}
suggested a bimodal distribution in redshift space, but were not at first considered significant.
Recent kinematic studies focussing solely on Abell~2744 definitely show a bimodal velocity dispersion
in the cluster centre, together with a third group of cluster members near the Northwestern X-ray peak \citep{Boschin2006,Braglia2009}.
Although there is no evidence for non-thermal X-ray emission \citep{Million2009}, the fraction of
blue star-forming galaxies \citep[][and references therein]{Braglia2007} also seems to be enhanced.
A default explanation emerged for the centre of Abell~2744, featuring a major merger in the North-South direction with a small
inclination towards the line-of-sight and a $\sim$~3:1 mass ratio of the merging entities \citep{Kempner2004,Boschin2006}.
More controversial is the role of the Northwestern structure.
\citet{Kempner2004} detected a cold front on its SW edge and a possible shock front
towards the cluster core, so concluded that it is falling towards the main mass.
More recent analysis of {\sl Chandra} data \citep{Owers2011} failed to confirm the presence of the shock
front, only a cold front towards the Northern edge, and the authors proposed that the structure is moving
instead towards the North/Northeast, after being deflected from the main cluster in an off-centre core passage.
What seems sure is that we are at least observing a complicated merger between three separate bodies \citep{Braglia2007}.
Until now, Abell~2744 has been
better constrained from X-ray and kinematic studies than by gravitational lensing.
So far, \citet{Smail1997} detected a weak-lensing signal and strong-lensing features, followed by \citet{Allen1998} who
found a large discrepancy in the mass estimates for Abell~2744 from X-ray and strong-lensing reconstructions. Given the cluster's dynamical
state, this finding is perhaps no longer surprising since the merging would induce non-thermal support and elongation along the line-of-sight,
which increases the systematics from both methods.
The most recent weak-lensing analysis \citep{Cypriano2004} showed indications of substructure
in the reconstructed surface-mass density, but did not reach the resolution required for more quantitative statements.
Here, we present the results of an {\sl HST} imaging survey, aimed at clarifying the
evolutionary stage of this complex system.
This article is organised as follows. In Sec.~\ref{LENSING} we present our comprehensive lensing analysis, mainly based on
newly acquired data taken with the {\sl Advanced Camera for Surveys (ACS)} on the {\sl Hubble Space Telescope (HST)}.
In Sec.~\ref{XRAY} we describe a complementary X-ray analysis based on {\sl Chandra} data.
We discuss cosmological implications and an interpretation of the merging scenario
in Sec.~\ref{INTERPRETATION}, and we conclude in Sec.~\ref{CONCLUSIONS}. Throughout this paper we assume a cosmological
model with $\Omega_{\textrm{m}}=0.3$, $\Omega_{\Lambda}=0.7$ and $\textrm{h}=0.7$.
At the cluster's redshift $z=0.308$, one arcsecond corresponds to $4.536$~kpc.
\section{Lensing analysis}
\label{LENSING}
Our recently acquired, multiband {\sl HST/ACS} imaging enables us to significantly improve upon previous mass models of Abell~2744.
Using the parametric method described in Sec.~\ref{SLENSING}, we have identified strong gravitational
lensing of 11 background galaxies producing 34 multiple images around the Southern core,
with an Einstein radius of $r_{\textrm{E}} \sim 30\arcsec$ (see below).
These enable us to tightly constrain the position and shape of the core mass distribution.
No such multiple image systems are revealed around the N or NW clumps, immediately indicating that their masses are lower.
Our {\sl HST} images also yield $\sim62~\textrm{galaxies}/\textrm{arcmin}^{2}$ for weak lensing analysis
(after charge transfer inefficiency (CTI) corrections are performed, as described below). This enables detailed mass modelling throughout our {\sl HST} field of view.
To probe the cluster merger on even larger scales, we incorporate ground-based weak lensing measurements from {\sl VLT} and {\sl Subaru}.
We simultaneously fit all of these strong- and weak-lensing observations using our well-tested mass reconstruction algorithm \citep{Merten2009,Meneghetti2010a}
(which is similar to that used by \citet{Bradav2006} and \citet{Bradav2008} to map the \textit{Bullet Cluster}
and the \textit{Baby Bullet}). Importantly, we make no assumptions about mass tracing light in the combined analysis. It should
be noted though, that we rely on a parametric method, assuming that light traces mass, to identify multiple image systems. We account
for possible uncertainties in this identification, by resampling the set of less confident identifications in the errors analysis based on a
bootstrapping technique (see Sec. \ref{WLSL}).
Our analysis reveals {\em four} individual clumps of mass $\lower.5ex\hbox{$\; \buildrel > \over \sim \;$} 10^{14} M_{\odot}$ within a $250$ kpc radius.
Previous weak lensing analysis of {\sl VLT} images alone had resolved but a single broad mass clump \citep{Cypriano2004}.
Below we describe our datasets, analyses, and results in more detail. The central cluster field and a preview on the matter and gas
distribution is presented in Fig.~\ref{DMXOVERLAY}, which also shows the contours encircling different areas of
specific probability, that a mass peak is located at this specific location of the reconstructed field.
\begin{figure*}
\begin{center}
\includegraphics[width=.75\textwidth]{./FIG1_FEATURES_lowres.png}
\end{center}
\caption{The field of Abell~2744, with different interesting features. The false-colour background is provided by {\sl HST/ACS}
(the two pointings can be identified by the higher, bluely background noise level), {\sl VLT} and {\sl Subaru} images on a field
size of $240\arcsec\times 240\arcsec$ ($\sim 1.1$ Mpc on a side). Overlaid in cyan are the surface-mass density contours most concentrated in the 'core' area
and in magenta the more evenly-distributed X-ray luminosity contours.
The peak positions of the core, N, NW, and W clumps are indicated by the green likelihood contours, derived from the bootstrap samples.
Contours are 86\%, 61\%, and 37\% of the peak likelihood for each clump. (Assuming a Gaussian probability distribution,
these would correspond to 0.3, 1, and 2-sigma confidence contours.) Note that the NW clump most likely peaks in the area
indicated by NW1 (in 95\% of the bootstrap realisations), but peaks also at NW2 in 54\% of the bootstrap realisations.
The small red circles show the position of the local overdensities in the gas distribution,
associated with each individual dark matter clump. The white rulers show the separation between dark matter peaks and the bright
clusters galaxies and local gas peaks.}
\label{DMXOVERLAY}
\end{figure*}
\subsection{The HST/ACS dataset}
\label{ACSdata}
The {\sl HST} data consist of two pointings in Cycle 17 (data taken between Oct.~27-30 2009,
Proposal~ID:~11689, P.I.:~R.~Dupke)
with $\sim 50\%$ overlap between the pointings. The images were taken
with the {\sl ACS/WFC} camera using three different filters, F435W ($16.2~\textrm{ksec}$\footnote{Equally split between the two pointings}),
F606W ($13.3~\textrm{ksec}$) and F814W ($13.2~\textrm{ksec}$).
The {\sl HST/ACS} camera had been in orbit for eight years when the imaging was acquired.
During this time above the protection of the Earth's atmosphere, its CCD detectors had been
irreparably damaged by a bombardment of high energy particles.
During CCD readout, photoelectrons are transported to the readout amplifier through a silicon lattice.
Damage to this lattice creates charge traps that delay some electrons and spuriously trail the image -- in a way that
alters the shapes of galaxies more than the gravitational lensing signal that we are trying to measure.
To undo this trailing and correct the {\tt \_raw} images pixel-by-pixel, we used the detector readout model of \citet{Massey2010},
updated for device performance post Servicing Mission 4 by \citet{Massey2010a}. The corrected data was then reduced via the standard CALACS
pipeline \citep{Pavlovsky2006}, and stacked using multidrizzle \citep{Koekemoer2002}.
\subsection{Strong lensing}
\label{SLENSING}
We concentrate our strong-lensing analysis on the reduced {\sl ACS} images only.
Several strong-lensing features are immediately identifiable by eye on the combined three-band image (Fig.~\ref{ACS_RGB}).
To find additional multiple images across the field of view, we apply the well-tested approach of \citet{Zitrin2009} \citep[see also][]{Zitrin2011b} to
lens modelling, which has previously uncovered large numbers of multiply-lensed galaxies in {\sl ACS}
images of Abell 1689, Cl0024, 12 high-$z$ MACS clusters, MS1358 and Abell 383 \citep[respectively,][]{Broadhurst2005,
Zitrin2009,Zitrin2011,Zitrin2010a,Zitrin2011a,Zitrin2011c} .
In the \citet{Zitrin2009} method, the large-scale distribution of cluster mass is approximated by assigning a
power-law mass profile to each galaxy, the
sum of which is then smoothed. The
degree of smoothing ($S$) and the index of the power-law ($q$) are
the most important free parameters determining the mass profile. A
worthwhile improvement in fitting the location of the lensed images
is generally found by expanding to first order the gravitational
potential of this smooth component, equivalent to a coherent shear
describing the overall matter ellipticity, where the direction of the
shear and its amplitude are free parameters. This allows for some flexibility in
the relation between the distribution of dark matter and the distribution of
galaxies, which cannot be expected to trace each other in detail. The
total deflection field $\vec{\alpha}_T(\vec\theta)$, consists of the
galaxy component, $\vec{\alpha}_{gal}(\vec\theta)$, scaled by a
factor $K_{gal}$, the cluster dark matter component
$\vec\alpha_{DM}(\vec\theta)$, scaled by (1-$K_{gal}$), and the
external shear component $\vec\alpha_{ex}(\vec\theta)$:
\begin{equation}
\label{defTotAdd}
\vec\alpha_T(\vec\theta)= K_{gal} \vec{\alpha}_{gal}(\vec\theta)
+(1-K_{gal}) \vec\alpha_{DM}(\vec\theta)
+\vec\alpha_{ex}(\vec\theta),
\end{equation}
where the deflection field at position $\vec\theta_m$
due to the external shear,
$\vec{\alpha}_{ex}(\vec\theta_m)=(\alpha_{ex,x},\alpha_{ex,y})$,
is given by:
\begin{equation}
\label{shearsx}
\alpha_{ex,x}(\vec\theta_m)
= |\gamma| \cos(2\phi_{\gamma})\Delta x_m
+ |\gamma| \sin(2\phi_{\gamma})\Delta y_m,
\end{equation}
\begin{equation}
\label{shearsy}
\alpha_{ex,y}(\vec\theta_m)
= |\gamma| \sin(2\phi_{\gamma})\Delta x_m -
|\gamma| \cos(2\phi_{\gamma})\Delta y_m,
\end{equation}
where $(\Delta x_m,\Delta y_m)$ is the displacement vector of the
position $\vec\theta_m$ with respect to a fiducial reference position,
which we take as the lower-left pixel position $(1,1)$, and
$\phi_{\gamma}$ is the position angle of the spin-2 external
gravitational shear measured anti-clockwise from the $x$-axis. The
normalisation of the model and the relative scaling of the smooth dark matter
component versus the galaxy contribution brings the total number of
free parameters in the model to 6. This approach to strong lensing is sufficient
to accurately predict the locations and internal structure of multiple
images, since in practice the number of multiple images
readily exceeds the number of free parameters, which become fully constrained.
Two of the 6 free parameters, namely the galaxy power law index $q$
and the smoothing degree $S$, can be initially set to
reasonable values so that only 4 of the free parameters need to be fitted at first.
This sets a very reliable starting-point using obvious
systems. The mass distribution is therefore well-constrained and
uncovers many multiple-images that can be
iteratively incorporated into the model, by using their redshift
estimation, from photometry, spectroscopy or model prediction and location in the image-plane.
At each stage of the iteration, we use the model to lens the most obvious lensed
galaxies back to the source plane by subtracting the derived
deflection field, then relens the source plane to predict the
detailed appearance and location of additional counter images, which
may then be identified in the data by morphology, internal structure
and colour. We stress that multiple images found this way must be
accurately reproduced by our model and are not simply eyeball
``candidates'' requiring redshift verification. The best fit is
assessed by the minimum RMS uncertainty in the image plane
\begin{equation} \label{RMS}
RMS_{\textrm{images}}^{2}=\sum_{i} ((x_{i}^{'}-x_{i})^2 + (y_{i}^{'}-y_{i})^2) ~/ ~N_{\textrm{images}},
\end{equation}
where $x_{i}^{'}$ and $y_{i}^{'}$ are the locations given by the
model, $x_{i}$ and $y_{i}$ are the real image locations, and the
sum is over the total number of images $N_{\textrm{images}}$. The best-fit solution is unique
in this context, and the model uncertainty is determined by the
locations of predicted images in the image plane. Importantly, this
image-plane minimisation does not suffer from the well known bias
involved with source plane minimisation, where solutions are biased by
minimal scatter towards shallow mass profiles with correspondingly
higher magnification.
In Abell 2744 we have uncovered a total number of 34 multiple images belonging to 11 background sources.
We label the different systems in Fig.~\ref{ACS_RGB}, which also shows the critical curve of the cluster derived from
the strong lensing model. The model predicts an Einstein radius of $r_{\textrm{E}} = 30\arcsec\pm3\arcsec$.
With such a large number of clear multiple images we can constrain the inner
mass distribution very well, and we shall incorporate these constraints into our joint strong- and weak-lensing analysis described in Sec.~\ref{WLSL}.
\begin{figure}
\begin{center}
\includegraphics[width=.45\textwidth]{./FIG2_ACS_CCURVE_lowres.png}
\end{center}
\caption{A zoom into the innermost core region of the {\sl HST/ACS} images.
Shown as a continuous white line is the critical curve of the cluster as it is derived from the strong-lensing model.
It assumes a source redshift $z_{\textrm{s}}=2.0$. Also shown are the approximate positions of the
identified multiple-image systems as they are listed in Tab.~\ref{MSYSTEMS} with a varying two-colour scheme to avoid
confusion of close-by images. These systems were found by the method of
\citet{Zitrin2009} and are not simple optical identifications.
The visible field size is $\sim100\arcsec\times100\arcsec$, translating to $\sim~450$ kpc on a side.}
\label{ACS_RGB}
\end{figure}
\begin{table}
\caption{The multiple-image system of Abell~2744.}
\label{MSYSTEMS}
\begin{center}
\begin{tabular}{cccc}
\hline
Image-ID & $x$ & $y$& $z$ \\
($\langle\textrm{\tiny{source}}\rangle$.$\langle\textrm{\tiny{image}}\rangle\langle\textrm{\tiny{additional knot}}\rangle$)&(\arcsec)&(\arcsec)&\\
\hline
1.1&$-35.10$& $-13.55$&$2.0\pm0.3$\\
1.2&$-30.15$& $-23.95$&\\
1.3&$0.10$& $-35.40$&\\
1.11& $-33.60$& $-16.50$&\\
1.21& $-31.50$& $-21.55$&\\
1.31& $1.60$& $-35.85$&\\
\hline
2.1& $9.30$& $-11.50$&$2.0\pm0.3$\\
2.2& $-34.25$& $12.35$&\\
2.3& $2.80$& $1.05$&\\
(2.4)& $-0.50$& $-7.10$&\\
2.11& $11.55$& $-7.65$&\\
2.21& $-32.55$& $13.90$&\\
2.31& $5.55$& $3.10$&\\
(2.41)& $-0.05$& $-4.15$&\\
\hline
3.1& $-10.10$& $22.65$&$4.0\pm0.3$\\
3.2& $-7.40$& $22.95$&\\
(3.3)& $27.15$& $2.20$&\\
\hline
4.1& $-18.25$& $-9.00$&$3.5\pm0.3$\\
4.2& $-29.25$& $-5.30$&\\
4.3& $18.05$& $-31.60$&\\
\hline
5.1$^{*}$& $8.85$& $29.00$&$4.0\pm0.5$\\
5.2$^{*}$& $3.85$& $31.60$&\\
5.3$^{*}$& $19.65$& $19.30$&\\
\hline
6.1$^{*}$& $-38.15$& $-5.95$&$3.0\pm0.5$\\
6.2$^{*}$& $-24.25$& $-28.30$&\\
6.3$^{*}$& $-0.60$& $-33.15$&\\
\hline
7.1$^{*}$& $-37.35$& $-7.85$&$3.7\pm0.5$\\
7.2$^{*}$& $-27.85$& $-26.10$&\\
7.3$^{*}$& $5.10$& $-34.85$&\\
\hline
8.1& $-10.70$& $20.90$&$4.0\pm0.2$\\
8.2& $-8.00$& $21.40$&\\
(8.3)& $16.10$& $5.90$&\\
\hline
9.1$^{*}$& $-6.65$& $-18.45$&$3.0\pm0.5$\\
9.2$^{*}$& $-2.80$& $-21.90$&\\
(9.3)& $-43.20$& $10.80$&\\
\hline
10.1$^{*}$& $-6.65$& $-20.65$&$3.0\pm0.5$\\
10.2$^{*}$& $-3.55$& $-22.75$&\\
10.3$^{*}$& $-44.90$& $11.00$&\\
\hline
(11.1)& $-16.00$& $-13.30$&$3.0\pm0.5$\\
(11.2)& $-34.20$& $-4.65$&\\
(11.3)& $10.70$& $-31.55$&\\
\hline
\end{tabular}
\end{center}
\medskip
If the image ID is shown in brackets, the image is not confidently reproduced by the lensing model described in Sec.~\ref{SLENSING}.
The systems marked with an asterisk, were nicely reproduced by the parametric lensing model but we allow them to be resampled
by our nonparametric reconstruction technique to allow for uncertainties in the number of identified strong lensing features.
All other images define a `confident catalogue' of multiple-image systems and are used in our subsequent analysis.
The x-and y-coordinates are relative to the BCG position ($\alpha_{\textrm{J2000}} =3.58611^{\circ}$, $\delta_{\textrm{J2000}} =-30.40024^{\circ}$)
in arcseconds. Redshifts of each system and their respective error are derived from the model predictions.
\end{table}
\subsection{Weak lensing}
\label{WLENSING}
Several areas outside the cluster core are of special interest, due to the complicated structure of this merging cluster.
We measure the shapes of background galaxies to derive a weak-lensing signal and extend our mass reconstruction over a much larger field.
Since our {\sl HST} data cover only a limited field of view, we also include {\sl VLT} and {\sl Subaru} imaging in our weak lensing analysis.
Additional {\sl HST} pointings in the future are crucial for a full interpretation of this merger, as we shall discuss in the course of
this analysis.
\subsubsection{HST/ACS}
To select lensed background galaxies, we obtained photometric redshifts
based on our three filters using Bayesian photometric redshifts (BPZ) \citep{Ben'itez2000,Coe2006}.
Cluster ellipticals at $z \sim 0.3$ occupy a unique region in our colour-colour space,
enabling us to obtain better than expected results.
Of 118 galaxies with published spectroscopic redshifts
(Owers et al. 2011 and references therein; all $z < 0.7$) within our field of view,
99 yield confident photo-zs, accurate to $\Delta z \sim 0.06(1+z)$ RMS with no significant outliers.
Background galaxies were selected as those with confident photo-z $> 0.5$.
This cut successfully excludes all 5 foreground and 86 cluster galaxies in our spec-z sample with confident photo-z
(though one cluster galaxy was assigned a less confident photo-z $\sim$ 0.8).
However, given our limited photometric coverage, we do not rule out some contamination of our background
sample with foreground and/or cluster galaxies.
We measure the weak gravitational lensing signal in the F814W {\sl HST} exposures,
which are the deepest and contain the most galaxies at high redshift.
We measure their shapes, and correct them for convolution by the Point-Spread Function (PSF), using the `RRG' \citep{Rhodes2000} pipeline
developed for the {\sl HST} COSMOS survey \citep{Leauthaud2007,Massey2007b}.
The RRG method is particularly optimised for use on high resolution, space-based data.
Since {\sl HST} expands and contracts as it warms in the sun or passes through the shadow of the Earth, even this telescope does not have a constant PSF.
However, 97\% of the variation in its PSF can be accounted for by variation in its focal length \citep[][the separation between the primary and secondary mirrors]{Jee2007a}.
We therefore measure its focal length by matching the shapes of the $\sim 12$ bright stars in each pointing to models created by raytracing through the optical design \citep{Krist2003}.
This achieves a repeatable precision of $1\mu m$ in the determination of the focal length, and we construct a PSF model from all those stars observed during the 600-orbit COSMOS survey at a similar focus \citep{Rhodes2007}.
We finally use this PSF model to correct the shapes of galaxies, and to obtain estimates of the amount by which their light has been sheared.
The result is a catalogue of $1205$ galaxies with shear estimates, corresponding to a density of $\sim62~\textrm{galaxies}/\textrm{arcmin}^{2}$.
\subsubsection{VLT/FORS1}
The complementary {\sl VLT} data for our weak lensing analysis is identical to that included by \citet{Cypriano2004} in a study of 24 X-ray Abell clusters.
The total field of view is $6\farcm8$ on a side, centred on the brightest cluster galaxy (BCG) and significantly exceeding the coverage of our {\sl HST} imaging.
$V$, $R$ and $I$-band imaging was obtained with the {\sl FORS1} camera between April and July 2001, with exposure times of $330$ s in each filter.
The data were reduced with standard IRAF routines and, to maximise the depth, we perform weak-lensing shape analysis on the combined VRI image.
Seeing conditions were excellent, with a FWHM of stars in the combined VRI image of $0\farcs59$.
We measure galaxy shapes and perform PSF correction using the \texttt{IM2SHAPE} method \citep{Bridle2002}.
This involves a two-step process to first map the PSF variation across the observed field using stars, then to model each detected galaxy, perform PSF correction and recover its ellipticity.
To remove foreground contamination and unreliable shape measurements from our catalogue, we apply magnitude cuts plus additional rejection criteria \citep[see][]{Cypriano2004}.
Since that work, we have improved the efficiency of foreground galaxy removal and now keep a higher density of background galaxies in our shear catalogue.
The result is a catalogue of $912$ galaxies with shear estimates, or $\sim 20~\textrm{galaxies}/\textrm{arcmin}^{2}$.
\subsubsection{Subaru/SuprimeCam}
To extend the total field of view even further, especially in the Northern areas of the cluster field, we obtained $1.68$ ksec $i'$-band imaging data with {\sl Subaru/SuprimeCam} during Semester S08B.
The data were reduced following \citet{Okabe2008} and \citet{Okabe2010,Okabe2010a}.
Astrometric calibration was conducted by fitting the final stacked image with the 2MASS data point source catalogue; residual astrometric errors were less than the CCD pixel size.
Due to poor weather conditions, the seeing size is as large as $1\farcs28$.
We measure galaxy shapes and perform PSF correction using the \texttt{IMCAT} package \citep[provided by][]{Kaiser1995} in the same pipeline as \citet{Okabe2010,Okabe2010a} with some modifications following \citet{Erben2001}.
Background galaxies were selected in the range of $22<i'<26~{\rm ABmag}$ and
${\bar r}_h^*+\sigma_{r_h^*}\simeq3.4<r_h<6.0~{\rm pix}$,
where $r_h$ is the half-light radius, and ${\bar r}_h^*$ and
$\sigma_{r_h^*}$ are the median and standard error of
stellar half-light radii $r_{h^*}$, corresponding to the half median
width of the circularised PSF.
The density of background galaxies in our final shear catalogue is $\sim 15~\textrm{galaxies}/\textrm{arcmin}^{2}$.
This is 30-50\% of typical values from images obtained during normal weather conditions \citet{Okabe2010,Okabe2010a}.
\subsection{Combined lensing reconstruction}
\label{WLSL}
In order to combine the weak- and strong-lensing constraints in a consistent way, we use the joint lensing reconstruction algorithm
described in \citet{Merten2009} \citep[see also][for a similar approach]{Bradav2005,Bradav2009}. This method has been extensively
tested in \citet{Meneghetti2010a} and proved its capability to faithfully recover the cluster mass distribution over a broad range of scales.
Our joint mass reconstruction is nonparametric, in the sense that it neither makes any \textit{a priori} assumptions
about the cluster's underlying mass distribution nor does it need to trace any light-emitting component in the
observed field. However, as described in Sec.~\ref{SLENSING}, we use a parametric method to identify the multiple image systems
in the field.
We reconstruct the cluster's lensing potential (its gravitational potential projected onto the plane of the sky) $\psi$ by combining measurements
of the position of the critical line and the reduced shear.
To do this, we divide the observed field into an adaptive mesh, which discretises all observed and reconstructed quantities.
A statistical approach is chosen to combine our various measurements,
by defining a multi-component $\chi^{2}$-function that depends on the underlying lensing potential and a regularisation term
$R(\psi)$ to prevent the reconstruction overfitting noise \citep[see][]{Merten2009,Bradav2005}
\begin{equation}
\chi^{2}(\psi) = \chi^{2}_{\textrm{w}}(\psi)+\chi^{2}_{\textrm{s}}(\psi)+R(\psi).
\end{equation}
The weak-lensing term is defined by the expectation value of the complex reduced shear in each mesh position $\left\langle\varepsilon\right\rangle$, which is obtained
by averaging the measured ellipticities of all background galaxies within that grid cell
\begin{equation}
\chi^{2}_{\rm w}(\psi)=\left(
\left\langle\epsilon\right\rangle -\frac{Z(z)\gamma(\psi)}{1-Z(z)\kappa(\psi)}
\right)_{i}\mathcal{C}^{-1}_{ij}\left(
\left\langle\epsilon\right\rangle -\frac{Z(z)\gamma(\psi)}{1-Z(z)\kappa(\psi)}
\right)_{j} ,
\end{equation}
where $Z(z)$ is a cosmological weight factor as defined e.g.~in \citet{Bartelmann2001}, and $\gamma = \partial\partial\psi/2$
is the shear of the lens, with the two components expressed in complex notation.
$\kappa = \partial\partial^{*}\psi$ is the convergence, where the complex differential operator in
the plane is defined as $\partial := (\frac{\partial}{\partial\theta_{1}} +\mathsf{i}\frac{\partial}{\partial\theta_{2}})$, with $\theta_{1}$
and $\theta_{2}$ being the two angular coordinates in the sky. The indices $i,j$ indicate the discretisation of the input data and
the lens properties, where we have to take into account the full $\chi^{2}$-function because the averaging process of background
galaxies might result in an overlap of neighbouring mesh points, expressed by the covariance matrix $\mathcal{C}_{ij}$.
The strong-lensing term is defined as
\begin{equation}
\chi^{2}_{{\rm s}}(\psi)=\frac{\left(\det{\mathcal A}(\psi)\right)^{2}_{k}}{\sigma^{2}_{{\rm s}}}=
\frac{\left((1-Z(z)\kappa(\psi))^{2}-|Z(z)\gamma(\psi)|^{2}\right)^{2}_{k}}{\sigma^{2}_{{\rm s}}}\;,
\end{equation}
where the index $k$ labels all pixels in the reconstruction mesh, which are supposed to be part
of the critical curve within the uncertainties $\sigma_{\rm s}$, given by the pixel size of the grid.
At these points, the Jacobian determinant $\det{\mathcal A}(\psi)$ of the lens mapping must vanish.
We iterate towards a best-fitting lens potential by minimising the $\chi^{2}$-function at each mesh position
\begin{equation}
\frac{\partial\chi^{2}(\psi)}{\partial\psi_{l}}\stackrel{\textrm{!}}{=}0 \quad \mathsf{with} \quad l\in[0,...,N_{\mathsf{pix}}]\;.
\end{equation}
In practice, we achieve this by translating this operation into a linear system of equations and invoking a two-level iteration
scheme \citep[see][and references therein]{Merten2009}.
In this analysis, we use all strongly lensed multiple-image systems from the confident sample (Tab.~\ref{MSYSTEMS}) together with their derived redshifts, and
combine weak lensing shear catalogues from all three telescopes (see Sec.~\ref{WLENSING}) as described in Sec.~\ref{CMERGING}.
The combination of strong lensing data and the density of background galaxies allows for a reconstruction
on a mesh of $72\times72$ pixels in the central region (corresponding to a pixel-scale of $8.4\arcsec/\textrm{pix}$)
and $36\times36$ pixels in the outskirts of the field (corresponding to a pixel-scale of $16.7\arcsec/\textrm{pix}$).
Error estimates were produced by bootstrapping the redshift uncertainties of the strong-lensing constraints, and
resampling those multiple image systems in Tab.~\ref{MSYSTEMS}, which are marked with an asterisk. The result are 500 different bootstrap
realisations of the refined cluster core. 150 bootstrap realisations of the cluster outskirts were produced by
bootstrapping the combined ellipticity catalogues. The number of bootstrap realisations is mainly constrained by runtime considerations. All error estimates
are calculated from the scatter within the full bootstrap sample. If not stated differently, the given value reflects
68\% confidence level.
\subsubsection{Combining the catalogues}
\label{CMERGING}
In order to combine the three different catalogues of ellipticity measurements we used the following strategy. Clearly, the {\sl Subaru}
catalogue delivered the largest field-of-view but it was derived from only single-band imaging and under bad seeing conditions. Therefore,
we decided to limit its field size to a box of $600\arcsec \times 600\arcsec$ around the centre of the cluster and to cut out the central
$\sim 400\arcsec \times 400\arcsec$ part, which was sufficiently covered by {\sl HST/ACS} and {\sl VLT} exposures with better data quality.
As a result, {\sl Subaru} data only covers the outermost 200\arcsec on each side of the field.\\
One might argue that the combination of {\sl HST/ACS} and {\sl VLT} ellipticities in the innermost centre of the cluster is problematic,
but indeed the galaxy density is three times higher in the {\sl HST/ACS} field and clearly dominating the adaptive-averaging process
of the nonparametric reconstruction algorithm that we used for the further analysis (compare Sec.~\ref{WLSL} and \citet{Merten2009}).\\
Another issue is the possible double-counting of galaxies, so we cross-correlated both catalogues with a correlation radius of 2.5\arcsec
and found 160 double-count candidates over the full {\sl HST/ACS} field. Given the pixel size of the final reconstruction, this translates
to an insignificant average double-count probability of less than one galaxy per pixel.
Furthermore, the weighting scheme of the adaptive-averaging process was implemented such, that the highest weight of the {\sl VLT} ellipticities
was identical to the smallest weight of the {\sl HST/ACS} ellipticities. Finally, the errors in the ellipticity measurement
are treated in the joint reconstruction method in a purely statistical way by deriving the variance of the weight-averaged sample
of ellipticities in each reconstruction pixel.\\
To derive a physical surface mass density from the scaled lensing convergence one needs to know at least the mean redshift of the background
galaxy population that was used to produce the ellipticity catalogues.
Problems with the different depths of the fields and therefore
with the final mass analysis should not be a crucial issue for a relatively low-redshift cluster like Abell~2744. However, the redshift for
determining the surface-mass density from the reconstructed convergence in the overlap area has been calculated as the galaxy-density weighted average of
both, the {\sl VLT} and the {\sl HST/ACS} populations. The redshifts of the strong lensing features (compare Tab.~\ref{MSYSTEMS})
were included for the determination of the core mass and the respective redshifts of the {\sl VLT} and {\sl Subaru} source distribution
was used in the outskirts of the field.
In order to test the effect of dilution in the outskirts of the field, we increased
the ellipticity values for all Subaru background galaxies by 15\% and repeated the reconstruction. This test is necessary due to
the single-band {\sl Subaru} imaging. As it turns out, the difference in the reconstructed convergence is marginal since the
ellipticity values are already low with large scatter in this area of the field. However, the effect was included in the determination
of the error budget for the reconstructed total mass and mass profile.
\subsection{Reconstruction results}
\label{LENSING_REC}
We obtain a map of the lensing convergence across the field (proportional to the projected mass)
by applying the Laplacian operator to the lensing potential on the adaptively refined mesh (see Fig.~\ref{CONV}).
We find a total mass (by assuming $h=0.7$ here and further on) within a radius of $1.3$~Mpc around the Core of $M(r<1.3\textrm{Mpc})=1.8\pm0.4\times 10^{15}M_{\odot}$, which is
in good agreement with kinematically derived masses \citep{Boschin2006}.
A mass determination within a field of ($1300 \times 750$) kpc (to compare easily to the work on the {\it Bullet Cluster} of \citet{Bradav2006})
centred on the Core density peak
yields $M(1.73 \textrm{Mpc}^{2}) = 7.4 \pm 1.0\times 10^{14} M_{\odot}$, rendering Abell~2744 comparable in mass or slightly less massive
than the \textit{Bullet Cluster} \citep{Bradav2006}.
The overall radial convergence and mass profile can be found in Fig.~\ref{RPROFILE}.
Most interestingly, our gravitational lensing analysis resolves four distinct sub-structures, indicated in Fig.~\ref{CONV} by the white circles.
We label these substructures as Core, Northwestern (NW, later on dubbed as `dark'), Western (W, later on dubbed as `stripped') and Northern (N) structure.
Please note that there is some indication of a possible double peak in the NW structure, as can be seen in Fig.~\ref{DMXOVERLAY}
and also in the clearly visible extension of the structure towards the West in Fig.~\ref{CONV}.
The Core, NW and W clumps are clear detections in the surface-mass density distribution with $11\sigma$, $4.9\sigma$ and $3.8\sigma$
significance over the background level, respectively.
\footnote{The background level was estimated in the following way: From the total convergence, shown in Fig.~\ref{CONV}, a radial area
of $75\arcsec$ around the Core, $70\arcsec$ around NW and $55\arcsec$ around N and W was cut out. From the remaining field the
mean and variance in the convergence level was calculated.}
Somewhat fainter with $2.3\sigma$ significance is the N structure, but it clearly
coincides with a prominent X-ray substructure found by \citet{Owers2011} and is therefore included in the further analysis (compare Sec.~\ref{XRAY}).
The positions of the mass peaks and their local projected masses within $250$~kpc are listed in Tab.~\ref{STRUCTURES} and shown in
more detail in Fig.~\ref{DMXOVERLAY}.
The new {\sl HST/ACS} images thus allow a striking improvement in our map of the mass distribution and reveal
the distribution of dark matter sub-structure in great detail for the first time.
We will refer to the individual mass clumps resolved here in our discussion of the X-ray analysis below.
\begin{figure}
\begin{center}
\includegraphics[width = .48\textwidth]{./FIG3_SLWL.png}
\end{center}
\caption{The convergence map of the cluster field for a source redshift of $z_{\textrm{s}}=\infty$ and a field size of $600\arcsec\times600\arcsec$, translating to $\sim 2.7$ Mpc on a side.
The black contours start at $\kappa_{0}=0.14$ with a linear spacing of $\Delta\kappa=0.047$. The four white circles with labels indicate
identified sub-clumps and the radius within which their mass is calculated. The radius of all four circles is $55.4\arcsec\approx250$~kpc.}
\label{CONV}
\end{figure}
\begin{table}
\caption{Structures identified within our lensing reconstruction.}
\label{STRUCTURES}
\begin{center}
\begin{tabular}{cccc}
\hline
Name & $x$ & $y$&$M(r<250\textrm{kpc})$ \\
&(\arcsec)&(\arcsec)&($10^{14}~M_{\odot}$)\\
\hline
Core&$-13^{+11}_{-8}$&$-4^{+6}_{-13}$&$2.24\pm0.55$\\[1ex]
NW&$71^{+11}_{-10}$&$84^{+15}_{-7}$&$1.15\pm0.23$\\[1ex]
W&$177^{+23}_{-17}$&$-30^{+11}_{-15}$&$1.11\pm0.28$\\[1ex]
N&$34^{+23}_{-32}$&$170^{+10}_{-28}$&$0.86\pm0.22$\\
\hline
\end{tabular}
\end{center}
\medskip
The x-and y-coordinates are provided in arcseconds, relative to the BCG position
($\alpha_{\textrm{J2000}} =3.58611^{\circ}$, $\delta_{\textrm{J2000}} =-30.40024^{\circ}$). The
68\% confidence limits
on peak positions are derived from
500 bootstrap realisations for the Core peak, and from the pixel size of the coarse weak lensing mesh in the reconstruction outskirts for
the other mass peaks. Masses assume $h=0.7$ and their 68\% confidence limits are derived from bootstrap realisations as described in the text.
\end{table}
\begin{figure}
\begin{center}
\includegraphics[width = .5\textwidth]{./FIG4_RPROFILES.png}
\end{center}
\caption{Radial mass profiles of the cluster, derived from our lensing analysis. Shown in blue (referring to the left $y$-axis) is the convergence profile for a source redshift $z_{s}=\infty$.
The large uncertainty for small radii arises from uncertainty in the exact position of the cluster centre. The cumulative mass
as a function of radial distance from the cluster centre is shown in red (referring to the right $y$-axis). The black vertical line indicates the distance of the
outermost multiple images from the cluster centre.}
\label{RPROFILE}
\end{figure}
\section{X-ray analysis}
\label{XRAY}
The most difficult part of interpreting merging clusters is determining the geometric configuration of the collision,
such as its impact velocity, impact parameter and angle with respect to the plane of the sky \citep{Markevitch2002}.
For this purpose, X-ray data becomes a crucial addition to lensing measurements.
The location of any shock fronts are revealed in the temperature of the intracluster medium (ICM).
Velocities can be inferred from the density and temperature of intracluster gas (if the merger axis is near the plane of the sky),
or through direct Doppler measurements (if the merger axis is near the line of sight).
\subsection{Reduction of the Chandra data}
We reanalysed all existing {\sl Chandra} data of Abell~2744 (listed in \citet{Owers2011}), using \texttt{CIAO} 4.3 with the calibration database CALDB 4.4.2.
We cleaned the data using the standard procedure\footnote{http://cxc.harvard.edu/ciao/guides/acis\_data.html} and kept events with grades 0, 2, 3, 4 and 6.
We removed the {\sl ACIS} particle background as prescribed for `VFAINT' mode,
and applied gain map correction, together with Pulse Hight Amplitude (PHA) and pixel randomisation.
Point sources were identified and removed, and the sky background to be subtracted from
spectral fits was generated from Blank-Sky observations using the {\tt acis\_bkgrnd\_lookup} script.
We fit spectra using \texttt{XSPEC} V12.6.0q \citep{Arnaud1996} and
we adopt {\sc Vapec} thermal emission model from atomic data in the
companion Astrophysical Plasma Emission Database \citep{Smith2001}, allowing for the variation
of several individual elemental abundances during the spectral fittings.
Galactic photoelectric absorption is incorporated using the {\tt wabs} model \citep{Morrison1983}.
Spectral channels are grouped to have at least 20 counts/channel.
Energy ranges are restricted to 0.5--9.5 keV.
Metal abundances are quoted relative to the solar photospheric values of \citet{Anders1989},
and the spectral fitting parameter errors are 1-$\sigma$ unless stated otherwise.
There is a known reduction of quantum efficiency at energies below 1~keV
due to a build-up of molecular contaminants on the optical blocking filter
(or on the CCD chips)\footnote{http://cxc.harvard.edu/cal/Links/Acis/\\acis/Cal\_prods/qeDeg/index.html}.
To prevent this from affecting our measurement of low energy line abundances, hydrogen column density and overall gas temperature,
we fix the column density to the cluster's nominal value of 1.6$\times$10$^{20}$~cm$^{-2}$ \citep{Dickey1990}.
To be conservative, we present results from only those regions of the CCDs best-suited for velocity analysis ({\sl ACIS-I} pointings 8477 and 8557);
for an independent sanity check, we also repeat the analysis with {\sl ACIS-S} pointing 2212.
We exclude pointings 7712 and 7915 because the regions of interest cross multiple CCDs, and interchip gain fluctuations could introduce spurious systematic effects when measuring the redshifts from X-ray spectra.
Since spectral models are (weakly) degenerate with cluster redshift, it is common for simultaneous
fitting routines to get stuck in local $\chi^2$ minima before they reach the global best fit.
We circumvent this via an iterative approach.
We perform initial fits while varying the redshift values within reasonable ranges (via the command STEPPAR in \texttt{XSPEC}).
We then fix the redshift, and refit full spectral models to infer gas temperature, metal abundances and normalisations.
Subsequently, we use these best-fit values as inputs in a new fit with the redshift free to vary.
This provides a more reliable estimation of the error on the redshift measurement.
We repeat the process until the best-fit redshift no longer changes between iterations.
\subsection{Previous X-ray observations \& interpretations}
Abell~2744 shows an extremely disturbed X-ray morphology (compare Fig.~\ref{XPROFILE}).
With 25~ksec of {\sl Chandra} {\sl ACIS-S3} imaging, \citet{Kempner2004} decided it is
the aftermath of a N-S collision between similar mass proto-clusters, at Mach~$>2.6$.
They also found tentative evidence for a cold front in the detached nearby NW `interloper', which they guessed was falling into the main cluster.
\citet{Owers2011} obtained a further $101$~ksec of {\sl ACIS-I} imaging.
The deeper data revealed the `Southern minor remnant core' (SMRC) to be colder (T$_X\sim$~7.5~keV) than its surroundings,
with a high temperature region to the SE (T$_X>15$~keV) that they interpreted as a shock front.
They concluded the SMRC had been a low-mass bullet that has passed through the `Northern major remnant core' (NMRC), leaving central tidal debris (CTD).
The pressure ratio of $\sim3$:1 across the shock front corresponds to a sky-projected shock velocity of 2150 km/s (for an average temperature T$_X\sim$~8.6~keV).
\citet{Owers2011} reversed the \citet{Kempner2004} model of the interloper,
concluding that it came originally from the South, has already passed through the main body of
the host cluster with a large impact parameter, and is now climbing out towards the N-NW.
\citet{Owers2011} also obtained spectra of more than 1200 galaxies with the {\sl Anglo-Australian Telescope}
multi-fibre {\sl AAOmega} spectrograph, and confirmed the velocity bi-modality of the cluster galaxies.
One component of galaxies near the SMRC has a peculiar velocity of $2300$~km/s; a separate component near the NMRC
has a peculiar velocity of about $-1600$\~km/s and enhanced metal abundances ($\sim$~0.5 Solar).
Assuming that the Northern and Southern cores have the same velocities as the `apparently' associated galaxy populations,
they de-projected the velocity of this collision to Mach~3.31 or nearly 5000~km/s.
An interesting prediction of this scenario is that the intracluster gas should show a strong radial velocity gradient
of $\sim4000$~km/s or $\Delta~z\sim 0.014$ from North to South.
Since the intracluster medium is enriched with heavy elements, radial velocity measurements could be carried out directly
by measuring the Doppler shift of emission lines in the X-ray spectrum
\citep{Dupke2001a,Dupke2001,Andersson2004,Dupke2007} or through changes in
line broadening due to turbulence \citep{Inogamov2003,Sunyaev2003,Pawl2005}.
The former requires high photon counts within the spectral lines and excellent control of
instrumental gain but could, in principle, be measured with current X-ray spectrometers.
The latter requires very high spectral resolution that should become available through the
future {\sl ASTRO-H} and {\sl IXO} satellites \citep[however, see][]{Sanders2010a,Sanders2011}.
The existing data are not ideal for measurements of ICM velocity structure with high precision due to the variation
of gain expected from the analysis of multiple observations with different pointings and in different epochs. However,
the expected radial velocity gradient of 4000km/s, predicted from the optical work of \citet{Owers2011} is higher than the expected
interchip and intrachip gain variations, so that it becomes possible to test, even if just for consistency,
the proposed merger configuration, as we did in this work.
{\sl Chandra} has a good gain temporal stability \citep{Grant2001} and
we shall control for spatial variations by performing resolved spectroscopy
of multiple cluster regions using {\it the same CCD location} in different {\sl ACIS-I} pointings.
\subsection{Velocity measurements}
We perform our velocity analysis twice: for the regions of interest defined by \citet{Owers2011}, then for the substructures prominent in our lensing mass maps.
These are respectively indicated by green or blue circles in Fig.~\ref{XPROFILE}.
We find temperature and metal abundance values for the NMRC, CTD and SMRC regions of
8.29$\pm$0.67 keV and 0.55$\pm$0.18 Solar ($\chi_{\nu}=$1.06 for 234 degrees of freedom);
9.51$\pm$0.47 keV and 0.27$\pm$0.08 Solar ($\chi_{\nu}=$1.03 for 469 degrees of freedom); and
7.98$\pm$0.76 keV and 1.17$\pm$0.51 Solar ($\chi_{\nu}=$1.01 for 143 degrees of freedom).
These values are derived from simultaneous fits to pointings 8477, 8557 and 2212.
The values are consistent with those obtained through individual spectral fits for each instrument/observation (two independent ACIS-I
and one ACIS-S). Our temperature and
abundance measurements are thus consistent with the values found by \citet{Owers2011}.
Given that the observations were not tailored for velocity measurements
(\textit{i.e.}\ we cannot exclude temporal or inter-chip gain variations),
we conservatively include in our velocity error bars secondary $\chi^2$ minima that are separated from the global minimum
at less than 90\% confidence in the velocity measurements.
Our velocity measurements are consistent with the presence of a gradient between the NMRC and SMRC,
but in the opposite sense to that expected from \citet{Owers2011} interpretation.
Using the ACIS-S, the CCD with the best spectral response, the velocity difference between NMRC and SMRC is found to be $>1500$~km/s at 90\% confidence.
The observations (8477 and 8557) indicate a higher magnitude for the velocity gradient detected. However,
the central values of the best-fit redshifts in each observation are found to be significantly discrepant, introducing larger uncertainties.
The reason for this discrepancy is not clear, but it is possibly related to the observed focal plane temperature variation of about 0.9C between
these observations (Catherine Grant Personal Communication).
This explanation is supported by the marginally significant observed systematic difference of $\sim$ 10\%-20\% in the best-fit values of temperatures and abundances measured for the same regions in these different ACIS-I observations. This effect was also noted in \citet{Owers2011} (Owers, M. Personal Communication).
Therefore, although our analysis is consistent with a high velocity gradient, the direction
of the gradient does not corroborate the merger configuration suggested by \citet{Owers2011}.
Having the lensing mass reconstruction shown in the previous sections at hand we can study the
line-of-sight gas velocity mapping for the regions of interest (Core and NW) more precisely.
Using all three exposures, we measure their gas temperatures and metal abundances to be
T$_{X_{Core}}$=10.50$\pm$0.57 keV, A$_{Core}$=0.28$\pm$0.11~Solar with a reduced chi-squared $\chi_{\nu}$=0.90 for 466 degrees of freedom and
T$_{X_{NW}}$=10.22$\pm$0.54 keV, A$_{NW}$=0.40$\pm$0.09~Solar with $\chi_{\nu}$=0.91 for 477 degrees of freedom.
Despite their similar gas properties, these regions show a velocity gradient $>5200$~km/s at near 90\% confidence,
even including a conservative (1$\sigma$) intra-chip gain variation of 1000~km/s \citep[e.g.][]{Dupke2006} per CCD in the error budget, in quadrature.
A contour plot of the velocity difference is shown in Fig.~\ref{XREDSHFITS}.
This result is consistent with the idea that the Southern Core mass is {\it redshifted} with respect to other structures.
Nonetheless, these calculations should be taken with caution due to the uncertainties related to temporal gain variations
and inter-chip gain variations between {\sl ACIS-I} and {\sl ACIS-S}.
The same analysis using only the {\sl ACIS-I} observations shows the same velocity trend,
but with a reduction in the absolute velocity difference (an upper limit of $<$4400~km/s at the 90\% confidence level).
Deeper {\sl Chandra} observations, specifically tailored for velocity measurements will be crucial to further reduce uncertainties
and to better constrain the velocity structure of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width = .48\textwidth]{./FIG5_CHANDRA.png}
\end{center}
\caption{X-ray image of Abell~2744, overlaid with our lensing mass reconstruction (magenta).
The velocity gradient is maximal between three regions of interest named by Owers et al. (2011)
and circled in green
(Southern minor remnant core, Northern major remnant core, and Central tidal debris).
A velocity gradient is also detected between the regions of interest from our lensing mass reconstruction, circled in blue
(Core and NW) as well as the interloper.
}
\label{XPROFILE}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width = .48\textwidth]{./FIG6_ZCONTOURS.png}
\end{center}
\caption{ Simultaneous fits to the redshifts of the Core region and the NW substructure using two {\sl ACIS-I} and one {\sl ACIS-S} pointings.
The two contours correspond to 68\% (inner) and 90\% (outer) confidence levels, and the diagonal line indicates equal redshifts to guide the eye. The Core mass clump appears significantly
redshifted when compared to the NW mass clump.
}
\label{XREDSHFITS}
\end{figure}
\section{Interpreting the merger}
\label{INTERPRETATION}
Abell~2744 is undoubtedly undergoing a complicated merging process on a large cluster scale.
Progressively more detailed studies, culminating in our lensing reconstruction (Sec.~\ref{LENSING}) and X-ray analysis (Sec.~\ref{XRAY}),
have only agreed that the merger is more complex than previously thought.
For example, the likely explanation for the gas velocity gradient in the opposite direction to that expected by \citet{Owers2011} is that the NMRC
is \textit{not} the main cluster.
Our lensing mass reconstruction shows that the deepest gravitational potential is by far the Southern `Core' structure,
which is roughly coincident with Owers et al.'s SMRC but slightly further offset from the Compact Core.
We also find three separate mass concentrations to the North, Northwest and West.
Overlaying the lensing mass reconstruction and X-ray emission in Fig.~\ref{DMXOVERLAY} reveals a complex picture of
separations between the dark matter and baryonic components.
To interpret the sequence of events that led up to this present state, we shall now tour the regions of interest, with more detailed discussions.
\subsection{Core, the massive clump}
According to our lensing analysis, the Core region (lower-left quadrant of Fig.~\ref{DMXOVERLAY}) is by far the most massive structure within the merging system (\textit{c.f.} Tab.~\ref{STRUCTURES}).
All the strong lensing features can be seen within this clump.
We find no large separation between the distribution of mass and baryonic components.
The mass peak is centred amongst the bright cluster member galaxies (within $1\sigma$ errors, it is consistent with the position of the BCG)
and only $22\pm12$ arcseconds from a peak of X-ray emission identified by \citet{Owers2011}.
We support the general conclusion of \citet{Owers2011} that the major-merger
in Abell~2744 is similar to that of the \textit{Bullet Cluster} as it would be seen at a large inclination with respect to the plane of the sky.
However, we reverse the ordering of the major and minor mass components.
One can infer constraints on the collisional cross-section of dark matter from the separation between peaks in the lensing and X-ray maps.
For the \textit{Bullet Cluster}, \citet{Markevitch2004} found $\sigma/m < 5~\textrm{cm}^{2}\textrm{g}^{-1}$, and
for the \textit{Baby Bullet}, \citet{ Bradav2008} found $\sigma/m < 4~\textrm{cm}^{2}\textrm{g}^{-1}$.
In the Core of Abell~2744, we observe a projected $17\arcsec$ separation that, if the inclination is
$\sim30^{\circ}$ away from the line-of-sight \citep{Owers2011}, is a physical separation similar to that in the other bullet clusters.
For an order of magnitude analysis, we measure the mean surface-mass density within $150$~kpc of the mass peak
$\Sigma\simeq0.30^{+0.08}_{-0.07}~\textrm{g}~\textrm{cm}^{-2}$, so that the scattering depth $\tau_{\textrm{s}} = \sigma /m~\Sigma$.
With the assumption $\tau_{s}<1$, which is justified due to the observed dark matter-gas separation, we deduce $\sigma/m < 3\pm1 ~\textrm{cm}^{2}\textrm{g}^{-1}$.
This system may therefore yield one of the tightest constraints on the interaction cross-section of dark matter, based on such analysis.
A full numerical simulation to interpret the cluster configuration would be ideal,
especially given the uncertainty in the collision angle with respect to the plane of the sky.
Indeed, even tighter constraints ($\sigma/m < 0.7 ~\textrm{cm}^{2}~\textrm{g}^{-1}$) were obtained from the \textit{Bullet Cluster}
by \citep{Randall2008}, who interpreted the offsets between all three cluster components via tailored hydrodynamical simulations.
Additional constraints on the collisional cross-section have also recently been obtained from
dark matter stripping in Abell~3827 \citep{Williams2011,Carrasco2010} and the ellipticities of dark matter halos \citep{Miralda-Escud'e2002}, with
implications discussed in \citet{Feng2010}.
\subsection{Northern, the bullet}
Our analysis of the Northern mass substructure (upper-left quadrant of Fig.~\ref{DMXOVERLAY}) confirms the overall North-South merging scenario proposed by several authors in the past.
We find a mass ratio of $\sim2.6$ between the Core and the Northern clump, roughly supporting the 3:1~merging scenario of \citet{Boschin2006},
but we identify the Northern sub-clump as the less massive progenitor.
This conclusion is robust, with no strong lensing features revealed by even our high resolution {\sl HST} imaging in the Northern structure,
as would have been expected for the reversed mass ordering proposed by \citet{Owers2011}.
X-ray emission in the Northern mass substructure lags behind the dark matter as expected.
We measure a separation of $\sim30\arcsec$ to the South.
This is a similar separation to that in the core but, due to the lower surface-mass density in this region, constraints on the collisional cross-section are less significant.
\subsection{Northwestern, the ghostly and dark clumps}
By far the most interesting structure is located to the Northwest of the cluster field (upper-right quadrant of Fig.~\ref{DMXOVERLAY}).
Our lensing analysis (Sec.~\ref{LENSING_REC}) shows it to be the second-most massive structure.
A separate region of X-ray emission also lies to the Northwest, called the NW interloper by \citet{Owers2011}.
Furthermore, in 54\% of our weak lensing bootstrap realisation we identified a second peak in the more Western area
of the NW mass clump (see Fig.~\ref{DMXOVERLAY}), rendering it difficult to say if we indeed see a single, separate dark matter structure
and to derive decisive separation between dark matter, X-ray luminous gas and bright cluster member galaxies.
However, the separation between our different possible NW mass peak positions and the NW interloper is large ($>150$~kpc for NW2).
When compared to the more prominent Eastern mass peak NW1, which is identified in 95\% of the bootstrap realisations, the distance
is even at least $400$~kpc.
With a de-projected temperature of $\sim5$~keV \citep{Owers2011} the NW interloper should have
$r_{500} \sim 1.34$~Mpc, \citep{Evrard1996} and $M_{500}\sim4-5\times10^{14} M_{\odot}$ (\textit{e.g.} Fig.~8 of \citet{Khosroshahi2007}).
Its $0.1-10.0$keV unabsorbed luminosity is $2.5\times10^{44}$~erg/s, consistent with its gas temperature of $4-5$ keV \citep{Khosroshahi2007,D'iaz-Gim'enez2011}.
Assuming a $\beta$ (from a King-like profile) of 0.67, typical for clusters, the clump would have $\sim 0.95\times10^{14} M_{\odot}$ within 250~kpc,
similar to the N clump, and should have been easily detected in the lensing analysis.
The interloper thus appears to be an X-ray feature with no associated dark matter or galaxies, and we therefore dub it the `ghost' cluster.
There is also a clear separation between the peak of the NW mass clump and any cluster member galaxies, so we call this the `dark' cluster.
Contours of the lensing mass reconstruction extend towards the West, where indeed we find a pair of giant ellipticals (see Fig.~\ref{DMXOVERLAY}).
However, with the limited resolution of the {\sl VLT} weak lensing reconstruction, it is impossible to tell whether this is a binary mass structure,
though there are clear indications for that in our analysis of this mass clump.
The separation between all three mass components makes this a real puzzle and it should be stressed that such a peculiar configuration
is observed for the first time. It may pose another serious challenge for cosmological models of structure formation.
One possible explanation was suggested by \citet{Owers2011}, who describe it as a ram-pressure slingshot \citep[e.g.][]{Markevitch2007}.
In this interpretation, after first core-passage, gas initially trails its associated dark matter but, while the dark matter slows down, the gas slingshots past it
due to a combination of low ram-pressure stripping and adiabatic expansion and cooling, which enhances the cold front temperature contrast \citep{Bialek2002}.
There is indeed a clear velocity gradient between the NW interloper and the main cluster core (see Sec.~\ref{XRAY}).
Such effects have also been observed e.g. in Abell~168 \citep{Hallman2004}, in a joint weak-lensing and X-ray study of Abell~754 \citep{Okabe2008}
and in numerical simulations \citep{Ascasibar2006,Mathis2005},
although at a much smaller separations between dark matter and gas.
The more than $100\arcsec$ separation in Abell~2744 suggests either that the slingshot scenario is unlikely or that some amplifying mechanism is in place.
We shall return to this issue, proposing an interpretation of the entire cluster merger, at the end of this section.
\subsection{Western, the stripped clump}
The Western substructure (lower-right quadrant of Fig.~\ref{DMXOVERLAY}) has not yet been discussed in the literature, but several cluster member galaxies are found in this area.
We find a prominent weak gravitational lensing signal of $\sim1.0\times10^{14} M_{\odot}$ within $250$~kpc.
This should correspond to an X-ray brightness higher than the NW interloper, for the same gas temperature.
However, we detect no X-ray emission.
The best X-ray data ({\sl ACIS-I} pointings 8477 \& 8557) do not cover the region of this lensing signal, but the {\sl ACIS-S} pointing 2212
indicates no excess diffuse gas above the cosmic background and the extended tail of the cluster's outskirts.
To match these observations, the Western clump must have been completely stripped of its ICM, so we dub it the `stripped' clump.
The only slight excess X-ray emission nearby is a faint extension of the main cluster core towards the West \citep[`ridge c' in][]{Owers2011}.
This roughly links the Western clump to the Northern clump,
and is consistent with remnants stripped by ram-pressure during a secondary merging event (NE--SW),
almost perpendicular to the main merging event projected in the sky plane.
To quantify the excess X-ray emission, we measure counts inside three equal-size, non-overlapping regions
extending from the cluster core to a radial distance of $150\arcsec\approx680$~kpc, as shown in Fig.~\ref{WBOX}.
To prepare the data for this analysis, we divide the image by the exposure map following the standard procedure, remove hot pixels
and remove point sources using the {\tt dmfilth} routine in \texttt{CIAO}.
We find a marginal excess in the central box (1030$\pm$32 counts), which points to the Western clump,
above the Northern (965$\pm$31 counts) and Southern (876$\pm$29 counts) boxes.
The cluster's surface brightness profile from the core towards the Western clump (along the central box) is shown in Fig.~\ref{WPROFILE}.
The slope of the profile changes at a radius $\sim135\arcsec=612$~kpc away from the centre, becoming significantly shallower.
If we take that point of transition as the location of the remaining gas core, we obtain a separation between the gas and dark matter of 30$\arcsec$.
Similar results are found with the shallower observation 7915 using {\sl ACIS-I}.
We find a slight offset between the peak of the lensing mass reconstruction and the most luminous nearby cluster member galaxies.
However, there is large uncertainty in the position of the lensing peak because this lies outside the {\sl HST} imaging area.
Our weak lensing analysis uses only {\sl VLT} imaging, and there are no strong lensing constraints, so the mass reconstruction has a broad central plateau.
Additional {\sl HST} observations would provide an ideal foundation to better understand the Western area,
which turns out to be playing a significant role in the overall merger.
\begin{figure}
\begin{center}
\includegraphics[width = .47\textwidth]{./FIG7_WBOXES.png}
\end{center}
\caption{{\sl Chandra} {\sl ACIS-S3} X-ray image of Abell~2744, smoothed with a 3 pixel kernel Gaussian.
We analyse the cluster's X-ray profile in three rectangular regions, each $150\arcsec$ in length and radiating from the cluster centre (marked by the blue circle).
We find marginally significant excess X-ray emission in the central rectangular region, which extends towards the Western clump (marked by the green circle).}
\label{WBOX}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width = .5\textwidth]{./FIG8_WPROFILE.png}
\end{center}
\caption{Surface brightness profile (in arbitrary units) of the central rectangular region shown in Fig.~\ref{WBOX}.
The left hand side starts at the X-ray centre, and the arrow denotes the approximate position of the Western mass clump.
The abrupt change in profile slope at roughly the same location is highlighted by the intersection with the horizontal line.}
\label{WPROFILE}
\end{figure}
\subsection{One possible interpretation}
We shall now try to develop a possible explanation of the complex merging scenario that has taken place in Abell~2744.
To recap, we find four mass clumps (Core, N, NW, W) with approximate masses 2.2, 0.8, 1.1, $1.1\times10^{14} M_{\odot}$, respectively.
The Core, N and W clumps are relatively close to BCGs and hot gas.
The NW structure, on the other hand, contains separated dark matter, gas and galaxies.
\begin{figure}
\begin{center}
\includegraphics[width = .5\textwidth]{./FIG9_MERGER_lowres.png}
\end{center}
\caption{Our proposed merging scenario, illustrated in time-ordered sequence, to explain the current configuration (panel 4).
We suggest that Abell~2744 is the result of a nearly simultaneous double merger: one in the NE-SW direction
and another in the NW-SE direction, which may even have consisted of three separate structures falling along a filament. Blue colour
shall indicate the innermost dark matter cores of the clumps, where else their respective ICM is shown in red.}
\label{FOURPANEL}
\end{figure}
We propose that the current configuration is the result of a near simultaneous double merger, as illustrated in Fig.~\ref{FOURPANEL}.
The first merger, in the NE-SW direction, had a characteristic path of $208\arcsec$ (plane of the sky distance between N and W clumps)
or $\sim0.95$ Mpc (assuming no line of sight velocity component).
The Western clump probably passed closest through the main cluster, as it had its ICM ram-pressure stripped completely.
The second merger, in the SE-NW direction, had a characteristic path of $117\arcsec/\sin(27^{\circ})\approx1.17$ Mpc,
if we assume the inclination of that merger suggested by \citet{Owers2011} of $27^{\circ}$.
The mergers happened around 0.12--0.15 Gyr ago, with a characteristic velocity of $\sim4000$~km/s
as indicated by the galaxy velocity difference and the ICM gas velocity measurements.
It is possible that the merger in the SE-NW direction could even have consisted of three initial substructures:
the Core and two consecutive clumps (with a combined mass (within 250 kpc of each core) of $\sim1.2\times10^{14} M_{\odot}$) falling along a filament.
Those smaller clumps would be accelerated by the gravitational pull of the main cluster
(plus the Northern and Western clumps, which were merging perpendicularly).
The ram-pressure slingshot in these clumps could be enhanced by a combination of an initially stronger gravitational field
and perhaps a posterior reduction in ram-pressure due to the `puff-up' of the gas density due to the recent merger of Core+N+W,
similar to the density configuration of the main component in the \textit{Bullet Cluster}.
The combined effect would throw the gas component ahead of its associated dark matter, forming the `ghost' cluster, which is now the interloper.
The two dark matter clumps left behind would now form the possibly double-centred `dark' clump.
This unusual scenario fits the current observations. It explains the clear extension of the NW clump towards the West and
provides probably a mechanism to create such a huge separation between dark matter and gas in the NW area. However, the double peak
in the NW clump needs to be confirmed by an additional {\sl HST} pointing to cover its full area. The discovery of a second distinct
mass peak in the NW clump would support our merger scenario and also decrease the large measured separations between dark matter, gas and
bright cluster galaxies.
Also the ghostly interloper has not been covered by the current {\sl HST} observations and the Western clump falls between
chip gaps in the longer {\sl Chandra} exposures, so further observations will still be required.
Our suggestions will also eventually require verification via a set of well-tailored numerical simulations and have to be taken with some
caution at the current stage. Cosmological boxes have to be explored \citep[compare][]{Meneghetti2010,Meneghetti2011} in order to find similar configurations
and detailed, hydrodynamical simulations need to repeat and confirm our findings.
\section{Conclusions}
\label{CONCLUSIONS}
We present a detailed strong lensing, weak lensing and X-ray analysis of the merging cluster of galaxies Abell~2744.
Earlier studies \citep{Kempner2004,Boschin2006,Braglia2009,Owers2011} concluded that Abell~2744 is undergoing a
complicated merging event. We find that it is even more complex than previously thought, unleashing a variety of exciting effects.
We dub this merger therefore \textit{Pandora's cluster}.
Deep, three-band {\sl HST}-imaging reveals a variety of strong-lensing features in the core of the cluster.
From our comprehensive strong-lensing modelling of the central mass distribution,
we identify a total number of not less than 34 multiple images, in 11 multiple-image systems,
together with their respective redshifts.
The strong lens systems are listed in Tab.~\ref{MSYSTEMS} and
the finely resolved critical curve of the cluster core is shown in Fig.~\ref{ACS_RGB}.
The Einstein radius of the core is $r_{E}=30\arcsec\pm3\arcsec$.
We extended the strong lensing information with weak lensing measurements over the whole cluster system ($600\arcsec\times600\arcsec$, or $\sim 2.7$ Mpc on a side).
The shapes of background galaxies necessary for weak lensing reconstruction were obtained from a comprehensive
combination of our new {\sl HST/ACS} imaging, the {\sl VLT} data used in \citet{Cypriano2004}, and from additional {\sl Subaru} imaging.
Our combined strong and weak lensing mass reconstruction (Fig.~\ref{CONV}) resolves a complex structure, with at least four distinct peaks in the
local mass distribution.
The total mass of the cluster is $1.8\pm0.4\times10^{15}M_{\odot}$ within a radius of $1.3$~Mpc, rendering Abell~2744 similar in mass to the \textit{Bullet Cluster}.
{\sl Chandra} X-ray imaging also shows a complex arrangement of substructure.
There are at least four X-ray ridges departing from the X-ray peak, including a Northern (NMRC) and a Southern (SMRC) ridge \citep{Owers2011} .
Interestingly, none of these coincide with any of the mass clumps found in our lensing mass reconstruction.
Furthermore, the system also has a separate Northwestern X-ray feature with very low mass, also undetected in our lensing analysis.
Observations of the gas temperature in this Northwestern feature \citep{Kempner2004,Owers2011} show a cold front pointing N-NE
and a shock region ($T_X>15$~keV) in the Southern ridge, for a projected impact velocity of 2150 km/s.
That projected direction (SE-NW) therefore seems to define the primary merging event.
However, in contrast to the proposed merging scenario of \citet{Owers2011} , we find the
Southern ridge to be blueshifted with respect to the Northern ridge.
We also reveal that the Southern Core is $\sim2.6$ times more massive than the Northern sub-clump --
\citet{Owers2011} had expected this reversed --
and that the secondary mass peak is in the Northwest.
The mass ratio thus remains in agreement with general kinematical studies of \citet{Boschin2006}, but swaps the sense of the collision from \citet{Owers2011}.
With this new scenario in hand, we repeated the velocity gradient analysis using the `true' mass clumps (Core and Northwestern).
These regions exhibit similar gas temperatures and metal abundances, and show evidence for a $>5200$~km/s velocity gradient with the Core region redshifted.
These limits should be taken with caution, since the {\sl Chandra} observations were not tailored specifically for velocity studies
(they were obtained in different CCDs and at different epochs, so could be affected by variations in detector gain),
and there is also some evidence of unknown calibration uncertainty between the two {\sl ACIS-I} pointings taken of the same patch of sky.
However, our results are consistent with the main merger having a significant component along the line-of-sight,
with a magnitude and orientation consistent with that seen in the bi-modal distribution of galaxy velocities.
We also find evidence for a second merging event, simultaneously with or just before the main merger.
The second merger, along the perpendicular NE-SW axis, was between today's Northern and Western mass peaks.
We think these collided inside the extended halo of the core.
During this dramatic collision, gas in today's Northern `bullet' clump was partially stripped by ram pressure,
creating a characteristic separation between dark matter and baryons similar to that seen in the \textit{Bullet Cluster}.
The Western `stripped' clump fared worse: all of its gas was removed, strewn into the tidal debris of the Core
and a faint trail of excess X-ray emission towards its current location where we find just dark matter and galaxies.
The smaller merger may have enabled a curious effect in the main, SE-NW merger.
We postulate that gas in the Core was puffed up by the first collision, reducing ram-pressure stripping during the second.
We also suggest that the main merger could have included two separate subclumps incident along a filament from the SE.
The combined effect would be an enhancement of the `slingshot' effect proposed by \citet{Owers2011},
by which the subclumps' gas was accelerated ahead of their dark matter.
This would explain the very large $>35\arcsec$ observed separation between any gas
(`ghostly' clump or NW interloper), galaxies and dark matter (`dark' clump), as well as the double-peaked morphology of the `dark' clump.
However, this scenario needs further confirmation.
The interpretation for this spectacular merging system will benefit immensely from additional observations
and also from numerical simulations that can try to reproduce the new phenomenology shown in this cluster.
The Western and Northwestern clumps have not been covered by
{\sl HST} observations and {\sl Chandra} observations of the Western clump are shallow.
Wide-field {\sl Subaru} imaging in good seeing conditions and in several colours would also be useful to interpret the global environment.
Numerical simulations should be performed to confirm the enhancement of the baryonic slingshot by the
complicated merger configuration.
The example of the \textit{Bullet Cluster} has shown that the combination of complete lensing and X-ray observations
\citep{Markevitch2004,Bradav2006} with highly resolved hydrodynamical simulations \citep{Springel2007,Randall2008} is a particularly
powerful tool to understand the physics of merging clusters.
In future work, we shall attempt to repeat this analysis on Abell~2744.
The challenge laid down will be to explain the complicated phenomenology associated with this multiple merger, as well as to
better constrain the collisional cross-section of dark matter, which our rough calculation suggests must be $\sigma/m < 3\pm1~\textrm{cm}^{2}~\textrm{g}^{-1}$,
a tighter constraint than that from a similar analysis of the \textit{Bullet Cluster}, which does not involve numerical simulations.
Therefore, with our own numerical follow-ups we should be able to place similar or even better constraints than e.g. \citet{Randall2008} for
the \textit{Bullet Cluster}.
\section*{Additional affiliations}
$^{11}$Department of Physics and Astronomy, 2130 Fulton St., University of San Francisco, San Francisco, CA 94117, USA\\
$^{12}$Department of Physics and Astronomy, University of British Columbia, 6224 Agricultural Road, Vancouver, BC V6T 1Z1, Canada\\
$^{13}$Instituto de Astrof\'{\i}sica de Andaluc\'{\i}a (CSIC), C/Camino Bajo de Hu\'{e}tor 24, Granada 18008, Spain\\
$^{14}$Department of Theoretical Physics, University of Basque Country UPV/EHU, Leioa, Spain \\
$^{15}$IKERBASQUE, Basque Foundation for Science, 48011 Bilbao, Spain \\
$^{16}$Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr, MS 169-327, Pasadena, CA 91109, USA \\
$^{17}$California Institute of Technology, 1201 East California Blvd, Pasadena, CA 91125, USA\\
$^{18}$Spitzer Science Center, MS 220-6, California Institute of Technology, Pasadena, CA 91125, USA\\
\section*{Acknowledgements}
The authors thank Matt Owers, Catherine Grant and Matthias Bartelmann for useful discussions.
We are particularly grateful to Douglas Clowe, the referee of this work, for his thorough reading of the manuscript.
His competent comments and suggestions
improved the quality of this work substantially.
JM acknowledges financial support from the Heidelberg Graduate School of Fundamental Physics (HGSFP)
and by contract research Galaxy clusters of the Baden-W\"{u}rttemberg Stiftung.
All runtime-expensive calculations were performed on
dedicated GPU-machines at the Osservatorio Astronomico di Bologna.
MM and JM acknowledge financial support from ASI (contracts I/064/08/0, I/009/10/0 and EUCLID-IC)
and from PRIN INAF 2009.
DC and RD acknowledge partial financial support from grant HST-GO-11689.09-A.
RD also acknowledges support from NASA Grant NNH10CD19C.
RM is supported by STFC Advanced Fellowship \#PP/E006450/1 and ERC grant MIRG-CT-208994.
ESC and LSJ acknowledge support from FAPESP (process ID 2009/07154) and CNPq.
NB and YJT acknowledge support from the Spanish MICINN grant AYA2010-22111-C03-00 and from the Junta de
Andaluc\'{\i}a Proyecto de Excelencia NBL2003.
This research was carried out in part at the Jet Propulsion Laboratory,
California Institute of Technology, under a contract with NASA.
\bibliographystyle{mn2e}
|
\section{Introduction}
The problem of option pricing \cite{bouch-pott,mant-stan,voit} has been a main
topic of investigation in much of the econophysics literature, challenged by the
well-known inadequacy of the standard Black-Scholes model to the real world
\cite{bouch-pott,mant-stan,voit, matacz,borland,lmori}. Options are an imperative
element in modern markets, since they play a fundamental role, as convincingly
shown long ago by Black and Scholes, in reducing portfolio risk. As an alternative to
the Black-Scholes model, one of the authors has implemented an option pricing scheme
which is based on the evaluation of statistical averages taken over samples generated
from the underlying asset log-return time series \cite{lmori}. This method, which we
will refer to as ``Empirical Option Pricing" (EOP), has been succesfully validated through
a careful study of FTSE100 options.
A deeper understanding of the statistical features of financial time series is in order, since
this would eventually allow us to replace real samples by accurate synthetic financial series,
improving the statistical ensembles used in EOP. As a closely related issue, our aim in this work
is to show that financial series can be hugely compressed (we mean lossy compression, in the
information theoretical sense) by wavelet-filtering, without spoiling option premium evaluation
by EOP. The low-pass wavelet-filtered signal contains log-return fluctuations defined on time
scales larger than a few hours and it is likely to yield, due to its high compression rate, a
more suitable ground for modeling and synthetization.
This paper is organized as follows. Secs. II and III provide brief accounts, respectively, of
the EOP method and of the low-pass wavelet-filtering procedure that has been applied to our
analysis of the FTSE100 index. The wavelet-filtered financial series, which keeps only $0.04 \%$
of the total number of wavelet components of the original signal is noted to encode the essential
statistical information needed for a consistent evaluation of FTSE100 option premiums with
expiration times larger than a single trading day. In Sec. IV, we summarize our findings and point
out directions of further research.
\section{Empirical Option Pricing (EOP)}
We rephrase here, without paying much attention to rigorous considerations, the main points
of EOP \cite{lmori}. Let $S(t)$ be an arbitrary financial index modeled as a continuous stochastic
process. More precisely, we write down a Langevin evolution equation for $S(t)$, which is a simple
generalization of the one underlying the Black-Scholes model \cite{hull, wilm_etal}:
\be
\frac{d S}{dt} = \mu(t) S + \sigma(t) \eta(t) S \ . \ \label{s_eq}
\ee
Above, $\mu(t)$ and $\sigma(t)$ are the time-dependent interest rate and volatility
of the index $S$. The stochasticity of the financial time series comes from the
gaussian white noise process $\eta(t)$ appearing in Eq. (\ref{s_eq}), which satisfies to
\bea
&&\langle \eta(t) \rangle = 0 \ , \ \nonumber \\
&&\langle \eta(t) \eta(t') \rangle = \delta(t-t') \ . \
\eea
Observe that both $\mu(t)$ and $\sigma(t)$ may be regarded as stochastic process
as well, with fluctuations correlated on time scales which are much larger than the
correlation time of $S(t)$.
Working within the It\^o prescription, Eq. (\ref{s_eq}) can be readily rewritten
as
\be
\frac{dx}{dt} = \sigma(t) \eta(t) \ , \ \label{dxdt}
\ee
where
\be
S(t) = S \exp \left [ \int_0^t dt'\tilde \mu(t') + x(t)-x(0) \right ] \ , \ \label{st}
\ee
with
\be
\tilde \mu(t) \equiv \mu(t) - \frac{1}{2} \sigma(t)^2 \ . \ \label{eq_mu-mu}
\ee
Above, $S \equiv S(0)$ is just the spot price of the index. We are interested, now,
to evaluate the premium of an european option which is negotiated with strike price $E$
and expiration time $T$. Similarly to what is done in the Black-Scholes treatment,
where $\mu(t)$ and $\sigma(t)$ are constant, the option premium $V$ (for, say, call options)
can be obtained from the computation of the statistical average
\be
V = \exp[-rT] \langle (S(T)-E) \Theta(S(T)-E) \rangle \ , \ \label{eq_v}
\ee
where $\mu(t)$ is replaced by $r$, the risk-free interest rate, in
the definition of $S(T)$ provided by Eqs. (\ref{st}) and (\ref{eq_mu-mu}).
For stochastic processes $\{x_n\}$ defined in discrete time, with time step $\epsilon$,
like real financial time series, Eq. (\ref{dxdt}) can be replaced by the finite difference
equation
\be
\frac{1}{\epsilon} (x_{n+1} - x_n) = \sigma_n \eta_n \ , \ \label{dxdt2}
\ee
where
\be
\eta_n \equiv \frac{\xi_n}{\sqrt{\epsilon}}
\ee
and $\xi_n = \pm 1$ is an arbitrary element of a discrete gaussian stochastic process,
defined by $\langle \eta_n \rangle =0$ and $\langle \eta_n \eta_m \rangle = \delta_{nm}$.
From (\ref{dxdt2}), we get, immediately,
\be
\sigma_n^2 = \frac{1}{\epsilon} (x_{n+1} - x_n)^2 \equiv \frac{1}{\epsilon}(\delta x_n)^2
\ee
and, therefore,
\be
\frac{1}{2} \int_0^T dt \sigma(t)^2 \simeq \frac{1}{2} \sum_{n=0}^{N-1}
(\delta x_n)^2 \ , \
\ee
where time instants are given by $t_n = n\epsilon$, with $T= N \epsilon$.
Substituting, now, (\ref{st}) and (\ref{eq_mu-mu}) (with $\mu(t)$ replaced by $r$)
in (\ref{eq_v}),
we get
\bea
S(T)&=& S \exp \left [ r T +x(T)-x(0) - \frac{1}{2} \sum_{n=0}^{N-1}
(\delta x_n)^2 \right ] \nonumber \\
&=& S \exp \left [ r T +\sum_{n=0}^{N-1} \left ( \delta x_n - \frac{1}{2} (\delta x_n)^2 \right ) \right ] \ . \ \label{st2}
\eea
It is important to note that $\delta x_n$, which appears in the above expressions
is, from Eq. (\ref{st}), nothing more than an element of the detrended log-return series,
i.e.,
\be
\delta x_n = \ln [S(t_{n+1})/S(t_n)] - \epsilon \tilde \mu(t_n) \label{dx} \ , \
\ee
where $\langle \delta x \rangle =0$, due to Eq. (\ref{dxdt}).
We are now ready to summarize EOP in the four following steps:
(i) A large period ($>$ two years) of reasonably statistically stationary high-frequency (minute-by-minute)
log-return series of the underlying asset is ``purified" by the remotion of outlier events (typically,
log-return fluctuations which are larger than 10 standard deviations) and of the mean one-week asset's
interest rate (detrending). The resulting series is a stochastic process $\{ \delta y_n \}$;
(ii) Since the historical volatility $\sigma = \sqrt{ \langle (\delta y_n)^2 \rangle }$ is in general different from
the volatility of the financial series during the option lifetime $T$, we introduce a correction factor $g$ to
define the stochastic process $\{\delta x_n = g \delta y_n\}$, which yields a putative volatility
$\sigma^* = g \sigma$ for that period \cite{comment}. The $g$-factor is the only adjustable parameter in EOP, which
accounts for the distinction between the past and the future behaviors of the financial index $S(t)$.
(iii) An ensemble $\cal{E}$ of samples, each of length $T=N \epsilon$ ($\epsilon = 1$ minute) is defined from
one-hour translations of the initial sequence $\{\delta x_0$, $\delta x_1$, ..., $\delta x_{N-1} \}$. In other words,
\be
{\cal{E}} = \bigcup_{m, \Delta} \{ \delta x_{m \Delta},
\delta x_{1 + m \Delta},...,\delta x_{N -1+ m \Delta} \} \ , \ \label{E}
\ee
where $m \in \mathbb{N}$ and $\Delta = 60$;
(iv) Option premiums are computed from (\ref{eq_v}), (\ref{st2}) and (\ref{dx}),
with statistical averages taken over the ensemble $\cal{E}$, defined in (\ref{E}).
We note, furthermore, that the optimal value for the $g$-factor is found through the least squares method, devised
for the comparison between the market and modeled option premiums.
A good agreement has been attained between the market and EOP values in a detailed study of the FTSE100 index
options \cite{lmori}. The comparison data is reported in the MKT and OP columns of Table I.
The performance of EOP would benefit greatly from the use of synthetic financial series which would
enlarge the ensemble of samples $\cal{E}$. Thus, one may wonder, having modeling aims in mind, on what are
the relevant statistical facts hidden in the financial time series. The essential question we address is, accordingly,
whether the financial series be decomposed into relevant and irrelevant contributions, as far as option pricing
is concerned. In the next section, we recall some ideas on wavelet-filtering, which have been crucial in the
investigation of this issue.
\section{Wavelet-Filtering and EOP}
\begin{figure}[tb]\includegraphics[width=11.0cm, height=15.0cm]{histograms.eps}
\caption{Monolog plots of the histograms of log-return fluctuations for both the original signal
(empty circles) and the wavelet-filtered series (filled circles) taken for time horizons of (a) 100 minutes,
(b) 300 minutes, and (c) 600 minutes. We call attention to the non-gaussian profiles of theses
distributions.}
\label{}
\end{figure}
Log-return fluctuations are constantly affected by avalanches of market orders which have to do with
speculative trends, and are clearly time-localized events. These features render the financial time series
suggestively adequate for wavelet analysis.
Since there is no requirement of continuity for the log-return time series, we have chosen, due to
easiness of handling, to work with Haar wavelets \cite{walker}. In the same way as for any other
discrete wavelet basis, the Haar wavelets are labelled by two integer indices $1 \leq j \leq J$ and
$0 \leq k \leq 2^j-1$ and are given by
\be
\psi_{jk}(t) = \psi_{00}(2^jt-k) \ , \
\ee
where
\be
\psi_{00}(t) = \left\{ \begin{array}{rcl}
1 & \mbox{for} & 0 \leq t < \frac{1}{2} \\
-1 & \mbox{for} & \frac{1}{2} \leq t < 1
\end{array}\right.
\ee
is the function known as ``mother wavelet". Observe that the above basis functions
are defined in the domain $0 \leq t < 1$.
\begin{center}
\begin{table*}[tb]
\begin{tabular}{|c|c|c|c||c|c|c||c|c|c|}
\hline
{Strike}
&\multicolumn{3}{c||}{02dec05 ($S=5528.1$) }&\multicolumn{3}{c||}{06dec05 ($S=5538.8$)}&\multicolumn{3}{c|}
{09dec05 ($S=5517.4$)}\\
\cline{2-10}
{Price}
&MKT&OP$$& ${\overline{\hbox{OP}}}$& MKT&OP&${\overline{\hbox{OP}}}$&MKT&OP& ${\overline{\hbox{OP}}}$ \\\hline
$5125$ & 410.5 & 412.67 & 413.00 & X & X & X & X & X & X \\
$5225$ & 312 & 312.79 & 313.12 & 324 & 321.87 & 322.25 & 298 & 297.51 & 297.74 \\
$5325$ & 214.5 & 213.94 & 214.19 & 225.5 & 222.87 & 223.15 & 199 & 197.72 & 197.87 \\
$5425$ & 122.5 & 121.93 & 122.17 & 131.5 & 129.48 & 129.65 & 103.5 & 102.35 & 102.20 \\
$5525$ & 50 & 48.61 & 48.55 & 53.5 & 53.52 & 53.41 & 29.5 & 29.72 & 29.15 \\
$5625$ & 13 & 13.01 & 13.00 & 12.5 & 14.97 & 14.90 & 3.5 & 4.79 & 4.65 \\
$5725$ & 2.5 & [0.60] & 0.60 & 2 & 1.66 & 1.65 & 0.5 & 0.35 & 0.29 \\
$5825$ & 0.5 & [0.0] & 0.0 & X & X & X & X & X & X \\
\hline
\hline
{Strike}
&\multicolumn{3}{c||}{19dec05 ($S=5539.8$)}&\multicolumn{3}{c||}{03jan06 ($S=5681.5$)}&\multicolumn{3}{c|}{12jan06 ($S=5735.1$)} \\
\cline{2-10}
Price&MKT&OP&${\overline{\hbox{OP}}}$& MKT&OP&${\overline{\hbox{OP}}}$&MKT& OP&${\overline{\hbox{OP}}}$ \\\hline
$5225$ & 329.5 & 332.38 & 333.05 & X & X & X & X & X & X \\
$5325$ & 234.5 & 237.49 & 237.64 & 368.5 & 368.36 & 368.84 & 414 & 414.96 & 415.12 \\
$5425$ & 148 & 149.46 & 150.09 & 271 & 268.86 & 269.30 & 314 & 315.02 & 315.18 \\
$5525$ & 76 & 77.75 & 77.97 & 177 & 176.67 & 174.97 & 215 & 215.09 & 215.25 \\
$5625$ & 28.5 & 30.91 & 31.09 & 93 & 91.20 & 91.40 & 119 & 116.81 & 116.81 \\
$5725$ & 8 & [5.18] & 5.09 & 34.5 & 34.40 & 34.44 & 40 & 34.72 & 34.38 \\
$5825$ & 2.5 & [0.60] & 0.54 & 9 & 8.41 & 8.39 & 5.5 & 4.51 & 4.36 \\
$5925$ & 0.5 & [0.0] & 0.0 & 2 & [0.19] & 0.18 & 0.5 & 0.20 & 0.16 \\
$6025$ & X & X & X & 0.5 & [0.0] & 0.0 & X & X & X \\
\hline
\end{tabular}
\caption{Call option premiums taken from the market (MKT) are listed together with the EOP evaluations performed with the original (OP) and
the wavelet-filtered (${\overline{\hbox{OP}}}$) series. The market data was recorded on 02dec05, 06dec05, 09dec05, 19dec05,
03dec06, and 12dec06; spot prices are indicated by $S$; the respective g-factors (and putative volatilities; see Sec. II) are $g=0.81$ ($\sigma^*= 6.1 \%$), $g = 0.91$ ($\sigma^* = 6.9 \%$), $g = 0.94$ ($\sigma^* = 7.1 \%$), $g=0.78$ ($\sigma^*= 5.9 \%$), $g = 0.83$ ($\sigma^* = 6.3 \%$), and
$g = 0.78$ ($\sigma^* = 5.9 \%$). The risk-free interest rate is $r=4.5\%$. The three first dates in december refer to options which expired
on 16dec05. For the other three dates, expiration date is 20jan06. The mean volatilities measured between 02dec05 and 09dec05 and between
19dec05 and 12jan06 were $\sigma = 8.0 \%$ and $\sigma = 6.1 \%$, respectively. The brackets in some of the OP$_0$ evaluations indicate
option premiums which are probably affected by poor sampling of the underlying financial series.}
\end{table*}
\end{center}
The detrended log-return series $\{ \delta x_0, \delta x_1, ..., \delta x_{N-1} \}$ of
length $N=2^{J+1}$ and zero mean \cite{comment2} can always be expanded in wavelet
modes as
\be
\delta x_i = \sum_{j=0}^J \sum_{k=0}^{2^j-1} c_{jk} \psi_{jk}(i/N) \ . \ \label{wav_exp}
\ee
Low-pass wavelet-filtering can be straightforwardly implemented from the expansion
(\ref{wav_exp}) by retaining the modes which have the scale index $j < j^\star$, where
$j^\star$ is an arbitrarily fixed threshold. We have taken (following the prescriptions
given in step (i) of EOP, as discussed in Sec. II) a financial time series of 241664 minutes
(around two years of data) for the the FTSE100 index, ending on 17th november, 2005.
The series is partitioned into 59 subseries, each of length 4096 (corresponding to $J=11$
and about two weeks of market activity), which are then wavelet-filtered with threshold
parameter $j^\star = 4$ (compression rate of $99.6 \%$).
Since wavelets with $j>0$ have zero mean, we expect that the histograms of log-returns
$ \sum_{n=0}^{N-1} \delta x_i$ will not be much affected for time horizons
$T = N \epsilon > 4096/2^{j^\star} = 256$ minutes. This is actually verified in Fig. 1. Therefore,
on the grounds of Eq. (\ref{st2}), it is clear that option prices can be alternatively estimated
through the use of the low-pass wavelet filtered series within EOP for time horizons which are
larger than one trading day ($T=510$ minutes).
In Table I, we report the computed premiums for call options based on the FTSE100 index with expiration
times ranging from a few days to one month, in december 2005 and january 2006. The agreement between
the original and the wavelet-filtered option premium evaluations is significative. It is important to recall,
as already indicated in Ref. \cite{lmori}, that the Black-Scholes framework is unable to yield
good estimates of the market option premiums listed in Table I.
\section{Conclusions}
We have found, taking the FTSE100 index as a case study, that its high frequency
(minute-by-minute) time series can be highly compressed for the purpose of option pricing.
The original and the low-pass wavelet-filtered series have remarkably similar performances
within EOP, even for a compression rate of $99.6\%$, which means that only 967 out of the
original 241664 wavelet components have been selected throught the wavelet-filtering
procedure. The retained wavelet coefficients have scale index $j$ smaller than the
fixed threshold $j^\star=4$ and are associated to log-return fluctuations defined on time
scales larger than a few hours. It turns out, thus, that one is entitled to use the filtered time
series to precify FTSE100 options with expiration times which are larger than just one trading day,
where log-returns are still clearly non-gaussian random variables. A promising approach to
option pricing, deserved for further investigation, is to address the problem of series synthetization
from the analysis of the statistical properties of the compressed wavelet-filtered financial indices
directly in wavelet space, in a spirit similar to what is done in the context of artificial multrifractal
series \cite{arneodo_etal2}. It is likely that EOP would, then, be considerably improved from
the use of much larger synthetic statistical ensembles.
\acknowledgements
This work has been partially supported by CNPq and FAPERJ.
The authors would also like to thank an anonimous referee for
a number of insightful comments, which have been of fundamental
importance in shaping the final form of this work.
|
\section{\bf Introduction }
The AdS/CFT
correspondence~\cite{Maldacena:1997re,Gubser:1998bc,Witten:1998qj}
provides a theoretical method to understand strongly coupled field
theories. Recently, it has been proposed that the AdS/CFT
correspondence can also be used to describe superconductor phase
transitions~\cite{Gubser:2008px,Hartnoll:2008vx}. Since the high
$T_c$ superconductors are shown to be in the strong coupling regime,
one expects that the holographic method could give some insights
into the pairing mechanism in the high $T_c$ superconductors.
There have been lots of works studying various holographic
superconductors, see review \cite{Hartnoll:2009sz,Herzog:2009xv}, in
which some effects such as scalar field mass, external magnetic
field, and back reaction {\it etc.}, have been discussed. Among
those works, some universality is discovered. For instance, the
critical exponents for the condensation near the critical
temperature is found to be $1/2$ which is a universal value in
mean-field theory \cite{Bardeen:1957mv}. However, this value of
$1/2$ is mostly obtained by numerically solving the holographic
systems and then fitting the data. The frequently asked question is
how to calculate this $1/2$ from analytical study. In
\cite{Siopsis:2010uq}, the authors used the variational method for
the Sturm-Liouville (S-L) eigenvalue problem\footnote{For short, we
will call this method used by \cite{Siopsis:2010uq} to be ``S-L
method" in the following context.} to analytically calculate some
properties of the holographic superconductors in a (2+1)-dimensional
boundary field theory. They managed to obtain the critical exponent
$1/2$ by perturbing the system near the critical temperature.
Afterwards, the authors in \cite{Zeng:2010zn} extended this S-L
method to study the holographic s, p, d-wave superconductors in
3-dimensional boundary field theory. They found that this $1/2$ is a
universal value for these models.
In this paper, we extended the S-L method to study holographic
p-wave and s-wave superconductors in a (4+1)-dimensional bulk with
Gauss-Bonnet gravity. It is not trivial to extend this analytical
studies to (4+1)-dimensional case, because the dimension of
space-time will affect the boundary behavior of the fields which
coupled to gravity. And moreover it would further have an impact on
the conformal dimension of the dual operators. For instance, in
(2+1)-dimensional boundary field theory the charge density
$\rho\propto T^2$ where $T$ is the temperature; However, in
(3+1)-dimensional boundary field theory the relation is $\rho\propto
T^3$. This discrepancy would make the analysis a little different
from the one in \cite{Siopsis:2010uq}. We will see it explicitly in
the following context. Besides, we also calculated the ratio between
the critical temperature and the charge density for various
Gauss-Bonnet couplings. The ratio was greatly consistent with the
previous numerical results \cite{Gregory:2009fj,Cai:2010cv}. We also
found that the greater the Gauss-Bonnet coefficients were, the
harder the condensations to form, which was also consistent with the
previous conclusions \cite{Pan:2009xa,Cai:2010zm,Brihaye:2010mr}.
Although the conformal dimensions of the dual operators on boundary
theory are different from those in \cite{Siopsis:2010uq}, there
still exists a universal critical exponents $1/2$ as in the
mean-field theory. The ratios between the condensation values and
the critical temperatures matched the numerical results up to the
same order.
The paper is organized as follows: In Sec.\eqref{sect:gb} we briefly
review the basic facts of Ricci flat Gauss-Bonnet-AdS black holes.
In Sec.\eqref{sect:pwave}, we will analytically calculate the ratios
between critical temperature and the charge density as well as the
critical exponents of holographic p-wave superconductors. The same
procedure will be followed in Sec.\eqref{sect:swave} by studying
holographic s-wave superconductors. We will draw a conclusion in
Sec.\eqref{sect:con}.
\section{\bf Ricci Flat Gauss-Bonnet-AdS Black Holes}
\label{sect:gb}
The action of Gauss-Bonnet gravity in a $5$-dimensional space-time
can be written as
\begin{equation}\label{action}
S=\int d^5 x \sqrt{-g} \Big(R+\frac{12}{L^2}+\frac{\alpha}{{2}}
(R^2-4R^{\mu\nu}
R_{\mu\nu}+R^{\mu\nu\rho\sigma}R_{\mu\nu\rho\sigma})\Big),
\end{equation}
where $12/L^2$ is related to the cosmological constant and $L$ is
the radius of the AdS space-time. The quadratic curvature term in
this action is the Gauss-Bonnet term with $\alpha$ the Gauss-Bonnet
coefficient. The Ricci flat solution of this action is the
$5$-dimensional Gauss-Bonnet-AdS black hole ~\cite{Cai:2001dz}. The
metric is described by
\begin{eqnarray}\label{metric}
ds^2
&=&-f(r)dt^2+\frac{1}{f(r)}dr^2+\frac{r^2}{L^2}(dx^2+dy^2+dz^2),
\\
f(r) &=&\frac{r^2}{2\alpha}\bigg(1-\sqrt{1-\frac{4\alpha}{L^2}(
1-\frac{ML^2}{r^4})}\bigg), \nonumber
\end{eqnarray}
where $M$ is the mass of the black hole. The horizon is located at
$r=r_+=\sqrt[4]{ML^2}$, and the temperature of the black hole is
\begin{equation}\label{temperatur}
T=\frac{r_+}{\pi L^2}.
\end{equation}
Here we should notice that in the asymptotic region with $r \to
\infty$,
\begin{equation}
f(r)\sim
\frac{r^2}{2\alpha}\Bigg[1-\sqrt{1-\frac{4\alpha}{L^2}}\Bigg].
\end{equation}
Therefore, one can define an effective radius $L_{\rm{eff}}$ of the
AdS spacetime by
\begin{equation}\label{L2}
L_{\rm{eff}}^2\equiv\frac{2\alpha}{1-\sqrt{1-\frac{4\alpha}{L^2}}}.
\end{equation}
We can see from this equation that one has to have $\alpha\leq
L^2/4$ in order to have a well-defined vacuum for the gravity
theory. The upper bound $\alpha =L^2/4$ is called Chern-Simons
limit. In the AdS/CFT correspondence, this asymptotic AdS spacetime
is dual to a conformal field theory living on the boundary
$r\rightarrow\infty$. The temperature of the black hole is just the
one of the dual field theory. For simplicity of making integration
and making comparison to the previous numerical results, we will
assume $0<\al\leq L^2/5$ in the following context.
\section{Holographic P-wave Superconductors}
\label{sect:pwave}
\subsection{Basic setup}
In order to study the holographic p-wave superconductors in the
probe limit, we will introduce an SU(2) Yang-Mills action in to the
bulk theory, the Lagrangian density is
\begin{equation}\label{action}
\mathscr{L}_p={-}\frac{1}{4}F^a_{\mu\nu}F^{a\mu\nu},
\end{equation}
where, $F^a_{\mu\nu}=\partial_\mu A^a_\nu-\partial_\nu A^a_\mu +
\epsilon^{abc}A^b_\mu A^c_\nu$ is the Yang-Mills field strength,
$a,b,c=(1,2,3)$ are the indices of the generators of SU(2)
algebra. $\mu,\nu=(t,r,x,y,z)$ are the labels of space-time with
$r$ being the radial coordinate of AdS. The $A^a_{\mu}$ are the
components of the mixed-valued gauge
fields
$A=A^a_{\mu}\tau^adx^{\mu}$, where $\tau^a$ are the SU(2)
generators with commutation
relation $[\tau^a,\tau^b]=\eps^{abc}\tau^c$.
$\epsilon^{abc}$ is the totally antisymmetric tensor
with $\epsilon^{123}=+1$.
Following Refs.~\cite{Gubser:2008wv,Manvelyan:2008sv,Basu:2009vv,Ammon:2009xh,Akhavan:2010bf}, we choose the ansatz of the gauge fields as
\be\label{ansatz} A(r)=\phi(r)\tau^3dt+\psi(r)\tau^1dx.\ee
In this ansatz we regard the U(1) symmetry generated by $\tau^3$ as the U(1) subgroup of
SU(2). We call this U(1) subgroup as U(1)$_3$. The
gauge boson with nonzero component $\psi(r)$ along $x$ direction is
charged under $A^3_t=\phi(r)$. According to AdS/CFT dictionary,
$\phi(r)$ is dual to the chemical potential in the boundary field
theory while $\psi(r)$ is dual to the $x$ component of some charged
vector operator $J$. The condensation of $\psi(r)$ will
spontaneously break the U(1)$_3$ gauge symmetry and induce the
phenomena of superconducting on the boundary field theory.
The Yang-Mills equations with the above ansatz (\ref{ansatz}) are
\be\label{ym} \left\{ \begin{array}{ll} \r_r^2\phi+\frac{3
}{r}\r_r\phi -\frac{L^2 \psi^2}{r^2 f}
\phi=0,\\
\\
\r_r^2\psi +(\frac{1}{r}+\frac{\r_rf }{f})\r_r\psi +\frac{\phi ^2
}{f^2}\psi=0.
\end{array}\right.
\ee In order to solve the equations (\ref{ym}), we need the boundary
conditions for the fields $\psi(r)$ and $\phi(r)$. At the black hole
horizon, it is required that $\phi(r_+)=0$ for the U(1) gauge field
to have a finite norm, and $\psi(r_+)$ should be finite. Near the
boundary of the bulk, we have
\begin{eqnarray}
\label{bphi}\phi(r)&\rightarrow& \mu - \rho/r^2 \\
\label{bpsi}\psi(r)&\rightarrow& \psi^{(0)} + \psi^{(2)}/r^2.
\end{eqnarray}
$\mu$ and $\rho$ are dual to the chemical potential and charge
density of the boundary CFT , $\psi^{(0)}$ and $\psi^{(2)}$ are dual
to the source and expectation value of the boundary operator $J_x^1$
respectively. We always set the source $\psi^{(0)}$ to zero, as we
want to have a normalizable solution.
\subsection{Relations between the critical temperature and charge density}
In this subsection, we will adopt S-L method, which was pioneered by
\cite{Siopsis:2010uq}, to analytically calculate the ratios between
the critical temperature $T_c$ and the charge density $\rho$ of the
holographic p-wave superconductors.
Define $z=\frac{r_+}{r}$, the EoMs \eqref{ym} will become
\be \label{ymz}
\left\{\begin{array}{c}
\phi''-\frac{\phi'}{z}-\frac{L^2\psi^2}{z^2 f}\phi=0,
\\\\
\psi''+(\frac1
z+\frac{f'}{f})\psi'+\frac{\phi^2r_+^2}{f^2z^4}\psi=0.
\end{array} \right.
\ee
where a prime denotes the derivative with respect to $z$. In the
following we will scale $L=1$. When the temperature approaches the
critical temperature {\it i.e.}, $T\rha T_c$, the condensation
approaches zero, viz. $\psi\rha0$, from eq.\eqref{ymz} we can get
\be \phi''-\frac{\phi'}{z}\approx0.\ee
With the boundary conditions \eqref{bphi}, we can get that near the
critical temperature, the electric field behaves like
\be\label{la} \phi(z)\approx \la r_+(1-z^2).\ee
where, $\la=\rho/r_+^3$. Near the boundary, we will introduce a
trial function $F(z)$ into $\psi(z)$ as in \cite{Siopsis:2010uq}
\be\label{J} \psi|_{z\rha0}\sim\psi^{(2)}\frac{z^2}{r_+^2}\sim \langle
J_x^1\rangle\frac{z^2}{r_+^2}F(z).\ee
where, $F(0)=1, F'(0)=0$. Therefore, using \eqref{la} and \eqref{J} the EoM of $\psi(z)$
reduces to
\be\label{Feom}
F''(z)&&+\frac{3+4\al(5z^4-3)-3\sqrt{1+4\al(z^4-1)}}{\big(1+4\al(z^4-1)-\sqrt{1+4\al(z^4-1)}\big)z}F'(z)
\nno\\&&+\frac{16\al
z^2}{1+4\al(z^4-1)-\sqrt{1+4\al(z^4-1)}}F(z)+\la^2\frac{4\al^2(z^2-1)^2}{(\sqrt{1+4\al(z^4-1)}-1)^2}F(z)=0.\nno\\\ee
Multiplying both sides of the above equation with
\be T(z)=\frac{z^3}{2\sqrt \al}\big(\sqrt{1+4\al(z^4-1)}-1\big),\ee
We will convert the eq.\eqref{Feom} to be
\be
\frac{d}{dz}\bigg(T(z)F'(z)\bigg)+\underbrace{\frac{8\sqrt{\al}z^5}{\sqrt{1+4\al(z^4-1)}}}_{-Q(z)}F(z)+\la^2
\underbrace{\bigg(\frac{4\al^2(z^2-1)^2T(z)}{(\sqrt{1+4\al(z^4-1)}-1)^2}\bigg)}_{P(z)}F(z)=0.\nno\\
\ee
From the Sturm-Liouville eigenvlaue problem \cite{Gelfand:1963}, the
minimum of eigenvalues of $\la^2$ can be obtained from the variation
of the following functional
\be \la^2[F(z)]=\frac{\int_0^1 dz\big(T(z)F'(z)^2+Q(z)F(z)^2\big)}{\int_0^1dz P(z)F(z)^2},\ee
The trial function $F(z)$ can be assumed to be
\be F(z)\equiv1-a z^2,\ee
which satisfies the boundary condition. So $\la^2(z)$ can be
explicitly written as
\be \la^2(z)=\frac{s(a,\al)}{t(a,\al)}.\ee
where,
\be s(a,\al)&=& 60(4 \alpha -1) \left((4 \alpha
-1) a^2+4 \alpha \right) \log \left(\frac{(1-4
\alpha ) \alpha }{\left(2 \alpha +\sqrt{\alpha
}\right)^2}\right) \nno\\&&-80 \sqrt{\alpha } \bigg(12
\alpha +a \left(20 \alpha a-3 a+32 \sqrt{1-4
\alpha } \alpha -48 \alpha -8 \sqrt{1-4 \alpha
}+8\right)\bigg),\nno\\
t(a,\al)&=&\frac{15}{2} (a+1) \alpha \left[128 (a+1) \log
\left(\frac{-4 \alpha +\sqrt{1-4 \alpha
}+1}{2-8 \alpha }\right) \alpha ^{3/2}\right.\nno\\&&\left.+\left(16
(3 a+4) \alpha ^2+8 (3 a+2) \alpha -a\right)
\log \left(\frac{(1-4 \alpha ) \alpha }{\left(2
\alpha +\sqrt{\alpha
}\right)^2}\right)\right]\nno\\&&-2 \sqrt{\alpha }
\bigg(\left(\alpha \left(32 \alpha \left(9
\sqrt{1-4 \alpha }-30 \log 2+5\right)+56
\sqrt{1-4 \alpha }-45\right)-2 \sqrt{1-4 \alpha
}+2\right) a^2\nno\\&&+5 \alpha \left(4 \alpha
\left(32 \sqrt{1-4 \alpha }-96 \log
2+11\right)+16 \sqrt{1-4 \alpha }-13\right)
a\nno\\&&+20 \alpha \left(\alpha \left(20 \sqrt{1-4
\alpha }-48 \log 2+6\right)+\sqrt{1-4 \alpha
}-1\right)\bigg)\ee
For different values of $\al$, we list the minimum value of $\la^2$
and the corresponding value of $a$ in Table\eqref{pwave}. Because
$T=r_{+}/\pi$ and $\la=\rho/r_+^3$, we can easily get that when
$T\sim T_c$, there is
\be T_c\approx\ga ~\sqrt[3]{\rho},\ee
where, $\ga=1/({\pi\la_{\rm min}^{1/3}})$.
In Table\eqref{pwave}, we also list the analytical values and
numerical values \cite{Cai:2010cv} of $\ga$ for different $\al$. The
differences between the analytical and numerical values are within
$1\%$. We can find that when $\al$ grows, the ratio $\ga$ will
decrease, which means the critical temperature will get lower
compared to the charge density. This implies that the greater the
Gauss-Bonnet coefficient is, the harder the condensation to form.
This statement is consistent with the previous numerical results
\cite{Gregory:2009fj,Cai:2010cv,Pan:2009xa,Cai:2010zm}.
\begin{table}
\caption{\label{pwave} The minimum values of $\la^2$ and the
corresponding $a$ for different $\al$ are listed in the first three
columns. The analytical and numerical values of $\ga$ and $\zeta$
are shown in the last four columns. The differences between the
analytical and numerical values of $\ga$ is about $1\%$ for various
$\al$'s. The order for $\zeta_{\rm analytical}$ and $\zeta_{\rm
numerical}$ is the same.}
\centering
\begin{tabular} {|c|c|c||c|c||c|c|}
\hline
$\al$ & $a$ & $\la_{\rm min}^2$ & $\ga_{\rm analytical}$ & $\ga_{\rm numerical}$ &$\zeta_{\rm analytical}$& $\zeta_{\rm numerical}$\\
\hline\hline
$0.0001$ & $0.685$ & $16.748$ & $0.199$ & $0.201$ &$326.858$&$499.036$\\
\hline
$0.1$ & $0.677$ & $20.604$ & $0.192$ & $0.194$&$337.735$&$509.796$ \\
\hline
$0.2$ & $0.653$ & $28.571$ & $0.182$ & $0.183$&$349.614$&$519.583$ \\
\hline
\end{tabular}
\end{table}
\subsection{Critical exponents and the condensation values}
From last subsection we know that when $T\rha T_c$, the condensation
value of the dual operator $\langle J_x^1\rangle$ is very small.
Using formula \eqref{J} we can convert the eq.\eqref{ymz} to be
\be\label{phieom}\phi''-\frac{\phi'}{z}=\frac{\mathcal{A}z^2F^2}{f}\phi,\ee
where, $\mathcal{A}=\langle J_x^1\rangle^2/r_+^4$. Because
$\mathcal{A}$ is small, considering \eqref{la} we can expand
$\phi(z)$ in $\mathcal{A}$ as
\be \frac\phi {r_+}\sim \la(1-z^2)+\mathcal{A}~\chi(z)+\cdots,\ee
From the boundary condition we can get $\chi(1)=0$, and for simplicity we set
$\chi'(1)=0$ \cite{Siopsis:2010uq}. The eq.\eqref{phieom} now becomes
\be \label{chieom}\chi''-\frac{\chi'}{z}=\la\frac{z^2F^2(1-z^2)}{f}.\ee
When temperature is away but very close to the critical
temperature, we expand $\phi(z)$ near $z=0$,
\be\label{expan}
\phi=\mu-\frac\rho{r_+^2}z^2=\frac{\rho}{r_+^2}(1-z^2)\Rightarrow
\frac\phi{r_+}&=&\frac\rho{r_+^3}(1-z^2)\sim\la(1-z^2)+\mathcal{A}\chi(z)+\cdots\nno\\
&=&\la(1-z^2)+\mathcal{A}\big(\chi(0)+\chi'(0)z+\frac{\chi''(0)}{2}z^2+\cdots\big).\nno\\\ee
Comparing the coefficients of $z^2$ term in the above formula~\footnote{Here, the analysis
is a little different from \cite{Siopsis:2010uq} in which they
compared the coefficients of $z$ term. The main difference comes
from the near boundary behavior of $\phi(z)$ which deeply originates
from the dimension of space-time. } we can get
\be \label{rho}\frac{\rho}{r_+^3}=\la-\frac {\mathcal A}{ 2}\chi''(0).\ee
From the eq.\eqref{chieom} we know that
\be
\chi''(0)=[\frac{\chi'(z)}{z}+\frac{\mathcal Az^2F^2}{f}\phi]\bigg|_{z\rha0}=[\frac{\chi'(z)}{z}]\bigg|_{z\rha0}\ee
Because the second term in the {\it r.h.s} of the first equality
approaches zero when $z\rha0$. In order to calculate the value of
$[\frac{\chi'(z)}{z}]\bigg|_{z\rha0}$, fortunately we find that
multiply $1/z$ to both sides of eq.\eqref{chieom} we get
\be(\frac1 z\chi')'=\la\frac{zF^2(1-z^2)}{f}\ee
Integrate both sides of the above equation and note that
$\chi'(1)=0$, we obtain~\footnote{Comparing the $z$ term in \eqref{expan}, we know that
$\chi'(0)\rha 0$ when $z\rha 0$. This is consistent with \eqref{chiprime}.}
\be \label{chiprime}[\frac{\chi'(z)}{z}]\bigg|_{z\rha0}=-\la\int_0^1dz
\frac{zF^2(1-z^2)}{f}\ee
Therefore, from \eqref{rho} one has
\be \frac{\rho}{r_+^3}=\la+\frac {\mathcal A}{ 2} \la\int_0^1dz
\frac{zF^2(1-z^2)}{f}\ee
From the form of $\mathcal A$ and $f(z)$ we can obtain
\be\frac{\rho}{\la r_+^3}=1+\frac{\mathcal B\langle J_x^1\rangle^2}{2
r_+^6}\ee
where $ \mathcal B=\int_0^1dz \big(zF^2(z)(1-z^2)/g(z)\big)$ and
$g(z)=f(z)/r_+^2$. So we can deduce that
\be \langle J_x^1\rangle=\zeta T_c^3\sqrt{1-\frac{T}{T_c}}.\ee
where $\zeta=\sqrt6\pi^3/\sqrt{\mathcal B}$.
The analytical values and numerical values \cite{Cai:2010cv} of
$\zeta$ for various $\al$'s can be found in Table.\eqref{pwave}.
They match up to the same order. Besides, in both analytical and numerical cases,
the ratio $\zeta$ increases when $\al$ grows bigger, which implies that the condensation
values for the dual operators will also increase. From the AdS/CFT dictionary, this
operator can be interpreted as the operator for the paring
mechanism, the bigger vacuum expectation value of this operator will
make the condensation harder. This statement is also in agreement
with the previous arguments \cite{Gregory:2009fj,Cai:2010cv,Pan:2009xa,Cai:2010zm}.
.
\section{Holographic S-wave Superconductors}
\label{sect:swave}
\subsection{Basic setup}
In order to study the holographic s-wave superconductors, we
consider the Maxwell field couples to a complex scalar field in the
bulk. The Lagrangian density is \cite{Hartnoll:2008vx},
\be \mathscr{L}_s=-\frac1
4F_{\mu\nu}F^{\mu\nu}-|\nabla\psi-iA\psi|^2-m^2|\psi|^2\ee
where $A_{\mu}(r)$ is the vector field and $m^2$ is the mass square of the
scalar field. As usual, we can choose the gauge of vector field as
$A_{\mu}=(\phi(r),0,0,0)$ which is consistent with the ansatz of
a real scalar field $\psi(r)=\psi(r)^*$. In the following, we also
scale the AdS radius to be $L=1$.
The EoMs for $\phi(r)$ and $\psi(r)$ are
\be\label{eom}
\left\{\begin{array}{c}
\r_r^2\phi+\frac3 r\r_r\phi-\frac{2\psi^2}{f}\phi=0, \\\\
\r_r^2\psi+(\frac{\r_r f}{f}+\frac3
r)\r_r\psi+(\frac{\phi^2}{f^2}-\frac{m^2}{f})\psi=0.
\end{array}\right.
\ee
Near the boundary, $\phi(r)$ and $\psi(r)$ behave like
\be\label{phibd} \phi(r)&=&\mu-\frac{\rho}{r^2},\\
\label{psibd} \psi(r)&=&\frac{\psi^-}{r^{\Dl_-}}+\frac{\psi^+}{r^{\Dl_+}}.\ee
where, $\mu$ and $\rho$ have the same interpretation as those in
Sec.\eqref{sect:pwave}. $\Dl_{\pm}=2\pm\sqrt{4+m^2L_{\rm eff}^2}$ is
the conformal dimension of the dual operator $J$ in the boundary
field theory. We will also set $\psi^-=0$ and $\langle
J\rangle=\psi^+$. Therefore, we will unambiguously define
$\Dl\equiv\Dl_+$ from now on. On the horizon, $\phi(r)=0$ because of
the well definition of the norm.
\subsection{Relations between the critical temperature and charge density}
Then, we can define $z=r_+/r$, the EoMs \eqref{eom} become
\be\label{eomz} \left\{
\begin{array}{c}
\phi''-\frac1 z\phi'-\frac{2\psi^2r_+^2}{fz^4}\phi=0,\\\\
\psi''+(\frac{f'}{f}-\frac1z)\psi'+(\frac{\phi^2}{f^2}-\frac{m^2}{f})\frac{r_+^2}{z^4}\psi=0.
\end{array}\right.
\ee
where a prime denotes the derivative with respect to $z$.
Because when $T\rha T_c$, the condensation is very small,
$\psi\sim0$. Therefore, the EoM of $\phi(z)$ becomes
\be \phi''-\frac1z\phi'=0.\ee
With the boundary condition \eqref{phibd}, we can deduce that near the critical temperature
$\phi(z)\sim\la r_+(1-z^2)$ where $\la\equiv\rho/r_+^3$.
As in the last section, we can introduce a trial function $F(z)$
into the asymptotical behavior of $\psi(z)$,
\be\psi\big|_{z\rha0}\sim\frac{\langle J\rangle}{r_+^{\Dl}}z^{\Dl}F(z).\ee
where $F(0)=1, F'(0)=0$.
Now, the EoM for scalar field reduces to
\be\label{Feoms} F''(z)+p(z)F'(z)+q(z)F(z)+\la^2 w(z) F(z)=0.\ee
where, \be p(z)&=&\frac{\left(1-\sqrt{4 \left(z^4-1\right) \alpha
+1}\right) (2 \Delta -3)+4 \alpha \left((2
\Delta -1) z^4-2 \Delta +3\right)}{z \left(4
\left(z^4-1\right) \alpha -\sqrt{4
\left(z^4-1\right) \alpha +1}+1\right)}\\
q(z)&=&\frac1 z\bigg[-\frac{\left(3 \left(\sqrt{4 \left(z^4-1\right)
\alpha +1}-1\right)-4 \left(z^4-3\right) \alpha
\right) \Delta }{-4 \left(z^4-1\right) \alpha
+\sqrt{4 \left(z^4-1\right) \alpha +1}-1}\nno\\&&+\frac{2 m^2 \alpha }{\sqrt{4 \left(z^4-1\right)
\alpha +1}-1}+(\Delta -1) \Delta\bigg] \\
w(z)&=&\frac{4 \left(z^2-1\right)^2 \alpha
^2}{\left(\sqrt{4 \left(z^4-1\right) \alpha
+1}-1\right)^2}.\ee
Multiply both sides of \eqref{Feoms} with
\be
T(z)=\frac{z^{2\Dl-3}}{2\sqrt{\al}}(\sqrt{1+4\al
(z^4-1)}-1),
\ee
we obtain the following equation
\be \bigg(T(z)F'(z)\bigg)'-Q(z)F(z)+\la^2P(z)F(z)=0\ee
where \be Q(z)&=&-T(z)q(z)\\
P(z)&=&T(z)w(z)\ee
Following the procedure used in the above section, we know that the
minimum value of $\la^2$ can be obtained from variation of the
following functional
\be \la^2[F(z)]=\frac{\int_0^1 dz\big(T(z)F'(z)^2+Q(z)F(z)^2\big)}{\int_0^1dz P(z)F(z)^2},\ee
It is easy to find that $F(z)=1-az^2$ and in the calculation we will
assume $m^2=-3/L_{\rm eff}^2$ in order to make comparison with the
previous results \cite{Gregory:2009fj}. So $\la^2(z)$ can be
explicitly written as
\be \la^2(z)=\frac{s(a,\al)}{t(a,\al)},\ee
where, \be s(a,\al)&=& -\frac{1}{768 \alpha ^2}\bigg[ 2
\sqrt{\alpha } \bigg(\left(96 \sqrt{1-4 \alpha }
\alpha +404 \alpha -75\right) a^2\nno\\&&+48 \left(14
\sqrt{1-4 \alpha } \alpha -24 \alpha -5 \sqrt{1-4
\alpha }+5\right) a+144 \left(2 \sqrt{1-4 \alpha }
\alpha +\alpha \right)\bigg)\nno\\&&+3 (4 \alpha -1)
\bigg(25 (4 \alpha -1) a^2+144 \alpha \bigg)
\sinh ^{-1}\left(2 \sqrt{\frac{\alpha }{1-4 \alpha
}}\right)\bigg]\\
t(a,\al)&=&\frac{ \sqrt{1-4 \alpha }}{480\al^{3/2}}\bigg(
\left(-144 \alpha ^2-28
\alpha +1\right) a^2-40 \alpha (8 \alpha +1) a-10
\alpha (20 \alpha +1)\bigg)\nno\\
&&+\frac{1}{2} (a+1)^2 \sqrt{\alpha } \log \left(-4
\alpha +\sqrt{1-4 \alpha }+1\right)\nno\\&&+\frac{(a+1) \left(16 \alpha (4 \alpha +1)+a \left(48
\alpha ^2+24 \alpha -1\right)\right) \log
\left(\sqrt{1-4 \alpha } \sqrt{\alpha
}\right)}{128 \alpha }\nno\\&&+\frac{1}{1920\al^{3/2}}\bigg[-960 (a+1)^2 \log (2-8 \alpha ) \alpha ^2\nno\\&&
-15 (a+1)
\left(16 \alpha (4 \alpha +1)+a \left(48 \alpha
^2+24 \alpha -1\right)\right) \log \left(2 \alpha
+\sqrt{\alpha }\right) \sqrt{\alpha }\nno\\&&+\left(920 \alpha ^2+90 \alpha -4\right) a^2+10 \alpha
(212 \alpha +13) a+960 (a+1)^2 \log (2) \alpha ^2\nno\\&&+40 (24 \alpha +1)
\alpha-960 (a+1)^2 \alpha ^2\Ga'(\frac3
2)/\Ga(\frac3 2) -960 (a+1)^2 \gamma_{\rm Euler} \alpha ^2\bigg]\ee
in which, $\Ga(z)$ is the {\it gamma function}
and $\gamma_{\rm Euler}$ is the {\it Euler's
constant}. Because $T=r_{+}/\pi$ and $\la=\rho/r_+^3$, we can easily
get that when $T\sim T_c$, there is
\be T_c=\ga ~\sqrt[3]{\rho},\ee
where, $\ga=1/({\pi\la_{\rm min}^{1/3}})$.
We list the analytical results in Table\eqref{tab}. We can find that
the values of $\ga$ obtained from the S-L approach are very close
to those calculated in \cite{Gregory:2009fj} by the matching
methods~\footnote{Consult Eq.~(3.27) in
Ref.\cite{Gregory:2009fj}.}. The differences are smaller than $4\%$
for various $\al$'s. We also find that $\ga$ decreases when $\al$
grows implying that the condensation is harder to form, which is
consistent with the previous results
\cite{Gregory:2009fj,Pan:2009xa}.
\begin{table}
\caption{\label{tab} For $m^2=-3/L_{\rm eff}^2$, the minimum values
of $\la^2$ and the corresponding $a$ for various $\al$'s are listed
in the first three columns. The analytical values from S-L approach
are very similar to those obtained from matching methods
\cite{Gregory:2009fj} with differences smaller than $4\%$. The
values of $\zeta$'s are matching up to the same order.} \centering
\begin{tabular}{|c|c|c||c|c||c|c|}
\hline
$\al$ & $a$ & $\la_{\rm min}^2$ & $\ga_{\rm S-L}$ & $\ga_{\rm matching}$&$\zeta_{\rm S-L}$&$\zeta_{\rm matching}$ \\
\hline\hline
$0.0001$&$0.722$ & $12.155$ & $0.210$ & $0.202$ &$7.705$&$6.369$ \\
\hline
$0.1$ & $0.706$& $14.290$ & $0.204$ & $0.197$ &$7.913$&$6.884$ \\
\hline
$0.2$ &$0.662$& $18.091$ & $0.196$ & $0.193$&$8.042$&$7.596$ \\
\hline
\end{tabular}
\end{table}
\subsection{Critical exponents and the condensation values}
When $T\rha T_c$, from the equation of $\phi(z)$, we get
\be\label{phiz2} \phi''-\frac1 z\phi'=\frac{2\mathcal A\phi z^{2\Dl-4}F^2}{g}\ee
where, $\mathcal A\equiv \langle
J\rangle^2/r_+^{2\Dl}$,$g(z)=f(z)/r_+^2$. Because near the critical
temperature $T_c$, $\mathcal A$ is small. We can expand $\phi(z)$
in $\mathcal A$ near $z=0$,
\be\label{phiz} \frac{\phi}{r_+}=\la(1-z^2)+\mathcal A\chi(z)+\cdots,\ee
Comparing the coefficients in $z^2$ terms, we can get from the above
formula that
\be\label{rhos} \frac{\rho}{r_+^3}=\la-\frac {\mathcal A}{ 2}\chi''(0)\ee
Substituting \eqref{phiz} into eq.\eqref{phiz2}, we can get EoM for
$\chi(z)$ as
\be \chi''-\frac1 z\chi'=2\la \frac{z^{2\Dl-4}F^2(1-z^2)}{g}\ee
Multiplying $1/z$ to both sides and following the procedure in last
section, we obtain
\be \chi''(0)=\frac{\chi'(z)}{z}\bigg|_{z\rha0}=-2\la\int_0^1dz
\frac{z^{2\Dl-5}F^2(1-z^2)}{g}.\ee
Therefore, \eqref{rhos} becomes
\be \frac{\rho}{r_+^3}=\la(1+\frac {\mathcal A}{
2}{\mathcal B})\ee
where $\mathcal B=2\int_0^1dz
z^{2\Dl-5}F^2(1-z^2)/g$.
Finally, using the formula of $\mathcal A$ we get that when $T\rha
T_c$, the condensation of the operator $J$ behaves like
\be \langle
J\rangle=\zeta (\pi T_c)^{\Dl}\sqrt{1-\frac T{T_c}}\ee
where, $\zeta=\sqrt6 /\sqrt{\mathcal B}$. In Table\eqref{tab} we list the
value of $\zeta$ in S-L method and those in matching method
\cite{Gregory:2009fj}\footnote{Refer to Eq.~(3.28) in Ref.\cite{Gregory:2009fj} }. We find that these two kinds of values are
matching up to the same order. The ratio $\zeta$ increases when
$\al$ grows bigger, which implies that the condensation will be
harder to form. This is consistent with the previous results~\cite{Gregory:2009fj,Pan:2009xa}.
\section{Conclusions}
\label{sect:con}
In this paper, we have used the S-L method to analytically calculate
the holographic p-wave and s-wave superconductors with Gauss-Bonnet
gravity in the probe limit. In particular, we calculate the ratios
$\ga$ between the critical temperature and the charge density on the
boundary. The ratios are in great agreement with the previous
numerical results or those obtained from the analytical matching
methods. We find that the ratio will decrease when the Gauss-Bonnet
couplings $\al$ increases, which implies that the bigger $\al$ is
the harder the condensation to form. This statement is in agreement
with the previous conclusions. Besides, we also calculate the ratio
$\zeta$ between the condensation value and the critical temperature.
Although the values of ratio $\zeta$ is not so well agreement with
the numerical results or those from matching methods, they are
consistent up to the same order. Moreover, we find that when $\al$
increases, the ratio $\zeta$ will also increase. This again implies
that the bigger $\al$ is the harder the condensation to form, which
also agrees with the previous conclusions. The universal critical
exponents $1/2$ in mean-field theory was also analytically confirmed
in both p-wave and s-wave case of Gauss-Bonnet gravity. This
critical exponents will not be affected by the dimension of the
space-time or the Gauss-Bonnet coefficients $\al$. This value $1/2$
is an exact property of the Lagrangian for the matter field of
systems, and has nothing to do with the gravitational
backgrounds~\cite{Liu}.
\acknowledgments
HFL and HQZ would like to thank Xin Gao, Bin Hu, Zhang-Yu Nie and
Yuan-Jiang Zhang for their helpful discussions and comments. HFL
would be very grateful for the hospitalities of the members in the
Institute of Theoretical Physics, Chinese Academy of Sciences. This
work was supported in part by the National Natural Science
Foundation of China (No. 10821504, No. 10975168, No.11035008 and
No.11075098), and in part by the Ministry of Science and Technology
of China under Grant No. 2010CB833004.
|
\section{Introduction}
Observations in the low-energy gamma-ray band from hundreds of keV to tens of MeV
provide a unique window on such phenomena as
gamma-ray lines from nuclear de-excitation produced by nucleosynthesis in supernovae, e.g., $^{56}$Ni, $^{56}$Co, and $^{44}$Ti
\citep{Matz_1988, Chevalier_1992, Ballmoos_1995, Schonfelder_2001, Cheng_2004, Boggs_2006};
gamma-ray lines from long-lived isotopes spread throughout our galaxy, such as $^{26}$Al and $^{60}$Fe \citep{Oberlack_1996, Knodlseder_1999, Diehl_2003};
electron--positron annihilation lines from the Galactic center \citep{Purcell_1997, Knodlseder_2003, Weidenspointner_2006};
neutron capture lines from solar flares \citep{Share_2000};
and gamma-ray resonant absorption along lines of sight toward gamma-ray bright quasars due to delta-resonance or giant dipole resonance \citep{Iyudin_2005}.
Furthermore,
radiation from black hole accretion disks has been detected \citep{McConnell_2002},
and radiation of neutral pions produced
by ions accelerated in the strong gravitational potential of black holes
is also expected to be detected \citep{McConnell_1996, Mahadevan_1997, Battacharyya_2003}.
We can also observe nonthermal processes in low-energy gamma-rays:
synchrotron radiation or inverse Compton-scattered gamma rays
in gamma-ray pulsars \citep{Aharonian_1998, Thompson_1999, Schonfelder_2000, Kuiper_2001},
active galactic nuclei \citep[AGN;][]{Urry_1995, Fossati_1998, Ghisellini_1998, Kubo_1998, Chiaberge_2001},
and gamma ray bursts \citep[GRB;][]{Briggs_1999, Paciesas_1999, Kaneko_2006}.
In addition,
there are diffuse galactic gamma rays probably produced by bremsstrahlung and inverse Compton scattering of electrons \citep{Boggs_2000, Strong_2000},
and diffuse extragalactic gamma rays, which are thought to be a combination of emissions from AGNs and Type Ia supernovae \citep{The_1993, Watanabe_1999}
or the combined emission from the Comptonization process
including nonthermal electrons in accretion disk coronae of AGNs \citep{Inoue_2008}.
In addition, the measurement of photon polarization through low-energy gamma-ray observations from nonthermal processes,
such as synchrotron radiation and Compton scattering,
would present a powerful diagnostic tool for astrophysics \citep{Lei_1997, Forot_2008}.
However, observation of this low-energy gamma-ray band is difficult because of the following reasons:
the dominant process in a detector is Compton scattering
and large backgrounds of photons are produced
in the hadronic process between cosmic rays and a satellite body.
Therefore, MeV gamma-ray astronomy has not advanced in comparison with X-ray or other gamma-ray bands.
In fact, COMPTEL onboard the Compton Gamma Ray Observatory ({\it CGRO})
discovered only $\sim30$ steady gamma-ray sources in the $0.75$--$30$ MeV band \citep{Schonfelder_2000},
whereas EGRET detected $\sim270$ sources \citep{Hartman_1999},
and {\it Fermi} found $1451$ sources during the first 11 months of the all-sky survey \citep{Abdo_2010}
in the sub-GeV/GeV region above $100$ MeV.
COMPTEL, which was the first Compton telescope onboard a satellite,
localized the direction of an incident gamma ray on an event circle superposed on the sky
by measuring the direction of a Compton-scattered gamma ray and the energies of both the scattered gamma ray and a Compton-recoil electron.
After the launch,
the sensitivity of COMPTEL was found to be mainly determined not by diffuse cosmic gamma rays
but by 1) locally produced background gamma rays;
2) non-Compton-scattering events such as multiple-photon events due to hadronic interaction, bremsstrahlung, or annihilation of positrons;
and 3) events due to other particles such as neutrons,
even after effective background rejection by time-of-flight measurement
between a Compton scattering material and an absorber of the scattered gamma rays \citep{Weidenspointner_2001}.
COMPTEL's limited sensitivity made it desirable to create a new instrument with a higher sensitivity in the sub-MeV/MeV region.
Therefore, we developed an electron-tracking Compton camera (ETCC)
consisting of a three-dimensional tracker of Compton-recoil electrons and an absorber of Compton-scattered gamma rays
by a new detection method with powerful background rejection \citep{Orito_2003, Tanimori_2004}.
Compared with COMPTEL, the ETCC can restrict the direction of an incident gamma ray to a reduced arc on the Compton circle
by measuring the three-dimensional track of the Compton-recoil electron
in addition to the energy of the Compton-recoil electron and the energy and direction of the Compton-scattered gamma ray.
Moreover, the angle between the direction of the scattered gamma ray and that of the Compton-recoil electron
can be used for powerful background rejection by checking the consistency between the measured and kinematically calculated values.
The ETCC can also detect the polarization of an incident gamma ray,
because the azimuth angle distribution of Compton scattering has an asymmetry for polarized gamma rays.
Because we had already confirmed the detection principle of the ETCC by ground-based experiments \citep{Takada_2005, Hattori_2007},
we initiated an experiment to observe sub-MeV gamma rays from celestial objects by using a balloon-borne camera,
the Sub-MeV gamma-ray Imaging Loaded-on-balloon Experiment (SMILE).
For the first flight of SMILE (hereafter SMILE-I),
in order to study the background gamma rays for observations of celestial objects and verify the background rejection capability,
we observed both
atmospheric gamma rays, generated by the interaction between cosmic-ray
particles and nuclei in the atmosphere, and diffuse cosmic gamma rays
at an altitude of up to $\sim 35$ km \citep{Takada_2007}.
Since the 1960s, several experiments have been conducted to observe diffuse cosmic gamma rays.
The first detection was using the {\it Ranger 3} satellite between $70$ keV and $1.2$ MeV in 1962 \citep{Metzger_1964}.
Thereafter,
the existence of these rays was confirmed using the satellites {\it ERS-18} between $0.25$ and $6$ MeV \citep{Vette_1970},
{\it OSO-3} between $7$ and $100$ keV \citep{Schwartz_1974},
{\it Apollo 15} in the energy range of $0.3$--$27$ MeV \citep{Trombka_1973},
and {\it Kosmos 461} between $28$ keV and $4.1$ MeV \citep{Mazets_1975},
and by using several balloons \citep{Bleeker_1970, Damle_1971, Makino_1975, Kinzer_1978}.
While these gamma-ray observatories used phoswich counters,
observations by using a balloon-borne double-Compton telescope in the energy range above $1$ MeV
began in the 1970s \citep{Schonfelder_1977, Schonfelder_1980}.
Subsequently,
observations by using satellites have continued with, for example,
the High Energy Astronomy Observatory \citep[{\it HEAO};][]{Marshall_1980, Kinzer_1997},
the Solar Maximum Mission \citep[{\it SMM};][]{Watanabe_1999},
{\it CGRO} \citep{Weidenspointner_2000, Kappadath_1996},
and {\it INTEGRAL} \citep{Churazov_2007} that is based on coded aperture imaging.
An MeV bump, which is in excess near a few MeV in the {\it Apollo} observations, remained a mystery for a long time but is now rejected,
because COMPTEL revealed
that it was an artifact from delayed gamma-rays produced in radioactive decays in the instrument \citep{Weidenspointner_2000}.
Atmospheric gamma rays have been observed by balloon experiments since the 1960s \citep{Peterson_1972, Kinzer_1974, Ryan_1977, Schonfelder_1980}.
Low-energy atmospheric gamma rays are mainly produced by bremsstrahlung from secondary cascade electrons \citep{Peterson_1973, Danjo_1972},
and the electron--positron annihilation line is also detected \citep{Peterson_1963, Ling_1977}.
Some models for atmospheric gamma rays exist, for example,
the Ling model \citep{Ling_1975} and the QinetiQ Atmospheric Radiation Model \citep[QARM;][]{Lei_2006}.
Previous observations showed
that the flux of atmospheric gamma rays depends on the zenith angle,
and models produced by the following scientists are based on the observational results:
Ling, Costa \citep{Costa_1984}, and Graser \citep{Graser_1977}.
Cosmic and atmospheric charged particles and neutrons could present large backgrounds in gamma-ray observations;
thus, the estimation of the fluxes of these particles is essential \citep{Wunderer_2006, Bowen_2007}.
The fluxes of these particles fluctuate on the basis of solar activity, observation altitude, and cutoff rigidity.
A few models for energy spectra of these particles at balloon altitudes have been proposed, such as Mizuno's model \citep{Mizuno_2004} and QARM.
Recently, in addition to an ETCC,
balloon experiments with new detection methods have been performed
to achieve higher sensitivity in the sub-MeV/MeV band.
For example, a double-Compton camera using a liquid Xe detector: the LXeGRIT \citep{Aprile_2008};
multiple Compton cameras \citep{Kamae_1987} by using semiconductors:
the nuclear Compton telescope \citep[NCT;][]{Boggs_2007}, a Si/CdTe Compton telescope \citep{Takahashi_2005},
and the Tracking and Imaging Gamma Ray Experiment \citep[TIGRE;][]{Zych_2008};
an electron-tracking Compton camera using an electron tracker made of a semiconductor:
the Medium Energy Gamma-ray Astronomy (MEGA) telescope \citep{Bloser_2006};
and a gamma-ray lens: CLAIRE \citep{Ballmoos_2005}.
Also, such experiments have renewed the attention given to observations of diffuse cosmic gamma rays, atmospheric gamma rays,
and the background formed by cosmic rays at balloon altitudes.
In this paper, we report the observational results of SMILE-I.
We first reconstructed the gamma-ray events and then
estimated the background produced by the interaction between cosmic ray and the materials surrounding the detector during the flight.
Next, we simulated the scattering of diffuse cosmic gamma rays in the atmosphere.
After that we obtained the fluxes of diffuse cosmic and atmospheric gamma rays measured with SMILE-I.
Finally, we determined the detection sensitivity of SMILE-I,
and estimated the expected detection sensitivity of a future SMILE mission.
\section{Instrument}
Figure~\ref{fig:mev_camera} shows a schematic of the ETCC developed by us.
The electron tracker detects the three-dimensional track and the energy of the Compton-recoil electron,
while the absorber detects the absorption point and energy of the Compton-scattered gamma ray.
By summing the momenta of both the recoil electron and the scattered gamma ray,
we obtain the momentum of the incident gamma ray as
\begin{equation}
p_0^\mu = p_\gamma^\mu + p_e^\mu ,
\end{equation}
where $p_0^\mu$, $p_\gamma^\mu$, and $p_e^\mu$ are the four-dimensional momenta of the incident gamma ray,
scattered gamma ray, and recoil electron, respectively.
Thus, we can obtain a completely ray-traced gamma-ray image.
The unit vector of the incident gamma ray ${\bf r}$ is described by
\begin{eqnarray}
{\bf r}
&= \left( \cos\phi - \frac{\sin\phi}{\tan \alpha} \right) {\bf g}
+ \frac{\sin\phi}{\sin\alpha} {\bf e} \\
&= \frac{E_\gamma}{E_\gamma + K_e} {\bf g}
+ \frac{\sqrt{K_e (K_e + 2 m_e c^2)}}{E_\gamma + K_e} {\bf e} \label{eq:kin_rcs},
\end{eqnarray}
where $\bf g$ and $\bf e$ are unit vectors in the directions of the scattered gamma ray and the recoil electron, respectively,
and $\phi$ is the scattering angle given by
\begin{equation}
\label{eq:phi}
\cos \phi = 1 - m_e c^2 \left( \frac{1}{E_\gamma} - \frac{1}{E_\gamma + K_e} \right) ,
\end{equation}
where $E_\gamma$, $K_e$, $m_e$, and $c$ are the energy of the scattered gamma ray, kinetic energy of the Compton-recoil electron,
electron mass, and light speed, respectively.
$\alpha$, which is the angle between the scattering direction and the recoil direction, as shown in Fig.~\ref{fig:mev_camera},
is geometrically measured for each incident gamma ray,
\begin{equation}
\cos \alpha_{geo} = {\bf g} \cdot {\bf e} ,
\end{equation}
and this angle is also obtained from a calculation by the Compton scattering kinematics:
\begin{equation}
\cos \alpha_{kin} = \left( 1 - \frac{m_e c^2}{E_\gamma} \right)
\sqrt{\frac{K_e}{K_e + 2 m_e c^2}} .
\end{equation}
By comparing $\alpha_{geo}$ with $\alpha_{kin}$,
\begin{equation}
\label{eq:alpha_cut}
\left| 1 - \frac{\cos \alpha_{geo}}{\cos \alpha_{kin}} \right| \leq \Delta_\alpha ,
\end{equation}
where $\Delta_\alpha$ is a cut parameter,
we can select only Compton scattering events.
Thus, Compton imaging with electron tracking is an effective method for MeV gamma-ray astronomy,
with the serious problem of a large background, as described in Section~1.
\begin{figure}
\plotone{fig1.eps}
\caption{
Schematic view of our electron-tracking Compton camera (ETCC).
The ETCC consists of a gaseous tracker that detects the Compton-recoil electrons
and a GSO:Ce pixel scintillator array (PSA) that detects the Compton-scattered gamma rays.
The distance between the drift plane and the gas electron multiplier (GEM) is $14$ cm,
and that between the GEM and the micro-pixel chamber ($\mu$-PIC) is $0.5$ cm.
The effective area of the time projection chamber (TPC) itself is the area between the drift plane and GEM.
\label{fig:mev_camera}
}
\end{figure}
For SMILE-I,
we have constructed an ETCC with a detection energy range $100$ keV--$1$ MeV.
Our ETCC consists of a gaseous tracker for detecting Compton-recoil electrons by using gas avalanche detectors to read it.
In addition, it has a pixel scintillator array (PSA) for detecting Compton-scattered gamma rays.
We provide a detailed description of our ETCC design in the following paragraphs.
As a tracker for detecting Compton-recoil electrons, we developed a gaseous time projection chamber \citep[TPC;][]{Kubo_2003, Miuchi_2003}
with a volume of $10 \times 10 \times 14$ cm$^3$ filled
with a gas mixture of $80$\% Xe, $18$\% Ar, and $2$\% C$_2$H$_6$ in mass ratio and sealed at $1$ atm.
The readout of the TPC consists of gas avalanche detectors: a gas electron multiplier \citep[GEM;][]{Sauli_1997, Tamagawa_2006}
and a micro-pixel chamber \citep[$\mu$-PIC;][]{Ochi_2001, Nagayoshi_D}.
The latter is our original gaseous two-dimensional imaging detector
with micro-pixel electrodes produced using printed circuit board technology.
A seed electron drifts into the $50$-$\mu$m-thick GEM with a velocity of $2.5$ cm $\mu$s$^{-1}$ at an electric field of $400$ V cm$^{-1}$,
and the first multiplication is caused by the GEM with a gain of approximately $10$.
Then, the multiplied electrons drift to the $\mu$-PIC,
and the second multiplication is caused by the $\mu$-PIC with a gain of $3 \times 10^3$.
Therefore, we obtain a high gas gain above $3 \times 10^4$, which is enough to detect Compton-recoil electrons.
The signals from the $\mu$-PIC are read using amplifier-shaper-discriminator chips \citep{Sasaki_1999},
which feed an output signal from a charge amplifier to a flash analog-to-digital converter to measure the recoil-electron energy,
and a discriminated digital signal to a position encoder
that encodes the electron track by using field-programmable gate arrays with a 100-MHz clock \citep{Kubo_2005}.
Because the $\mu$-PIC consists of fine-structure pixels with a pitch of $400$ $\mu$m,
this TPC has a fine three-dimensional position resolution of $\sqrt{\sigma_{\perp}^2 + \sigma_{\parallel}^2} = 500$ $\mu$m,
where $\sigma_\perp$ and $\sigma_{\parallel}$ are resolutions perpendicular and parallel to the drift direction, respectively \citep{Takada_2007}.
As an absorber for detecting Compton-scattered gamma rays, we used $33$-pixel-scintillator arrays \citep{Nishimura_2007},
each of which consists of $8 \times 8$ GSO:Ce scintillator pixels with a pixel size of $6 \times 6 \times 13$ mm$^3$.
For the photon sensor of the scintillation camera,
we selected a multi-anode photomultiplier tube (Hamamatsu Photonics, Flat-Panel H8500),
which consists of $8 \times 8$ anode pixels with a pixel size of $6 \times 6$ mm$^2$.
With chained resistors connecting the anodes to reduce the number of readout channels,
we obtained the position of a hit pixel by the charge-division method \citep{Sekiya_2006}.
A plastic scintillator with a size of $30 \times 30 \times 0.3$ cm$^3$ was placed $21$ cm above the GEM
as a veto counter for reducing triggers by charged particles.
When an energy deposit in the GSO scintillators was over $30$ keV,
a trigger was generated and the data-acquisition system waited for $8$ $\mu$s
(the maximum time for seed electrons to drift to the $\mu$-PIC) for the TPC signals.
When a trigger from the veto counter was generated,
the system waited for $100$ $\mu$s (the recovery time of the undershoot of the large signal caused by charged particles) after the veto trigger.
\begin{figure}
\plotone{fig2.eps}
\caption{
Cross-sectional view of a mass model of the SMILE-I ETCC.
\label{fig:mass_model_etcc}
}
\end{figure}
\begin{figure}
\plotone{fig3.eps}
\caption{
Dependence of detection efficiency on the energy of incident gamma rays
for a point source near the detector, obtained by experiments (circles) and simulation (triangles).
The dependence of efficiency on the incident energy for parallel light at a zero degree zenith angle
from the Geant4 simulation using a mass model of Fig.~\ref{fig:mass_model_etcc},
is also shown (solid line).
The error bars represent only statistical uncertainties.
\label{fig:efficiency}
}
\end{figure}
\begin{figure}
\plotone{fig4.eps}
\caption{
Effective area as a function of zenith angle
for a point source near the detector obtained by experiments (filled circles) and simulations (open circles),
and for parallel light with energies of 150 keV (dashed line), 600 keV (solid line), and 1500 keV (dotted line),
obtained by Geant4 simulation using a mass model shown in Fig.~\ref{fig:mass_model_etcc}.
The error bars represent only statistical uncertainties.
\label{fig:effective_area}
}
\end{figure}
To obtain the fluxes of diffuse cosmic gamma rays and atmospheric gamma rays by SMILE-I,
the detection efficiency and the effective area for parallel light is required.
However, irradiating the entire system of SMILE-I by using parallel light in a laboratory is very difficult.
Instead, we simulated the irradiation with Geant4 \citep[ver 9.0-patch01;][]{Agostinelli_2003} by using a mass model shown in Fig.~\ref{fig:mass_model_etcc}.
In the ETCC simulator,
the absorber had an energy resolution of $12 \times \left( {\rm Energy} / 662 {\rm keV} \right)^{-0.57}$ \%
at full width at half maximum (FWHM) and an energy threshold of $30$ keV,
the tracker had an energy resolution of $45 \times \left( {\rm Energy} / 32 {\rm keV} \right)^{-0.22}$ \%
at FWHM and an energy threshold of $\sim 3$ keV,
and the veto counter had an energy resolution of $50$\% at FWHM and an energy threshold of $300$ keV.
The time resolutions of the absorber and the veto counter were not considered,
because our ETCC did not use the event selection based on the time-of-flight between the Compton scattering target and the absorber.
We obtained the detection efficiency of gamma rays reconstructed
within zenith angles below $60$ degrees for irradiation from the zenith direction
as a function of the incident energy, as shown in Fig.~\ref{fig:efficiency};
the detection efficiency was approximately $10^{-4}$ between $150$ keV and $1.5$ MeV.
To estimate the systematic uncertainty of the detection efficiency,
we measured and also simulated the detection efficiency for a radioisotope
under the condition of gamma-ray irradiation from a point source placed at a distance of $50$ cm from the drift plane of the TPC.
The difference between the experimental measurement and simulated data was approximately $15$\%,
as shown in Fig.~\ref{fig:efficiency},
and we incorporated this difference into the systematic uncertainty for the detection efficiency.
In order to verify our mass model simulation of parallel light and to determine if the detection efficiency uncertainty depends on the zenith angle,
we plotted the effective area obtained by simulation for parallel light
between $150$ keV and $1.5$ MeV as a function of zenith angle (Fig.~\ref{fig:effective_area}).
For the verification of the simulator, we measured and also simulated the effective area for a point radioisotope.
The difference between the simulation and experimental data does not show any systematic dependence on incident direction.
Thus, we verified our mass model simulation, and assumed that the systematic uncertainty of the detection efficiency is not dependent on the zenith angle.
Figure~\ref{fig:effective_area} also shows
that the acceptance field-of-view (FOV) of SMILE-I was $120$ degrees ($3$ sr) full width at half maximum (FWHM) for $150$ keV,
$200$ degrees ($7$ sr) for $600$ keV, and more than $7$ sr for $1500$ keV, respectively,
which was wider than that of COMPTEL that was $2$ sr (FWHM) at $1.2$ MeV and $0.5$ sr at $6.1$ MeV \citep{Schonfelder_1993}.
Table~\ref{tab:res} shows the energy and the angular resolutions of the ETCC as a function of the incident energy,
measured at a ground-based experiment.
The angular resolution of the ETCC is defined by two parameters:
the angular resolution measure (ARM), which is the accuracy of the scattering angle $\phi$,
and the scatter plane deviation (SPD), which is the determination accuracy of the Compton-scattering plane \citep{Bloser_2002}.
In general,
the theoretical limit of the ARM is Doppler broadening caused by the undetectable momentum of the Compton-target electron,
and the SPD is limited by the multiple scattering of the Compton-recoil electron;
thus, an ETCC using a gas with a higher atomic number has a poorer angular resolution due to the larger multiple scattering,
although it has a higher detection efficiency.
The ARM resolution due to Doppler broadening is $3.3$ degrees for $200$ keV and $0.7$ degrees for $1$ MeV at FWHM.
The SPD resolution due to multiple scattering of the recoil electron with recoil energy of tens of keV is approximately $100$ degrees at FWHM.
In fact, an ETCC using an Ar gas had the ARM and SPD of $8.4$ degrees and $89$ degrees (FWHM), respectively, at $662$ keV \citep{Takada_2007},
and these values were better than those obtained using the higher atomic number Xe gas in the SMILE-I ETCC detector.
Because we intended to observe not a celestial point source
but diffuse cosmic gamma rays and atmospheric gamma rays in SMILE-I,
we gave preference to the higher efficiency over the better angular resolution and, thus, adopted Xe gas for our ETCC.
\begin{deluxetable}{cccc}
\tablecolumns{4}
\tablewidth{0pc}
\tablecaption{Energy resolution and angular resolutions (given by the FWHM) of the ETCC \label{tab:res}}
\tablehead{
\colhead{Energy} &\colhead{Energy resolution} &\colhead{ARM} &\colhead{SPD} \\
\colhead{[keV]} &\colhead{[\%]} &\colhead{[degrees]} &\colhead{[degrees]}
}
\startdata
166 &$37\pm2$ &$46\pm2$ &$189\pm17$ \\
356 &$20\pm2$ &$24\pm2$ &$181\pm10$ \\
511 &$15\pm1$ &$21\pm2$ &$180\pm15$ \\
662 &$14\pm1$ &$18\pm1$ &$183\pm13$ \\
835 &$15\pm2$ &$21\pm4$ &$185\pm17$
\enddata
\end{deluxetable}
For the balloon flight,
the ETCC was placed in an aluminum vessel with a diameter of $1$ m, a height of $1.4$ m, and a thickness of $3$ mm.
The vessel was maintained at $1$ atm
and the vessel was fixed to an aluminum gondola with a size of $1.2 \times 1.5 \times 1.6$ m$^3$.
On two sides of the gondola,
batteries and ballast boxes were attached, and the gondola was packed with expanded polystyrene.
In addition,
the SMILE-I gondola had a fine pressure gauge for measurement of atmospheric pressure,
a global positioning system receiver for measurement of the balloon's altitude and geographic position,
and two clinometers and two geomagnetic aspectmeters for determining attitude.
\section{Balloon Flight}
The SMILE-I balloon was launched from
the Sanriku Balloon Center of the Institute of Space and Astronautical Science/Japan Aerospace Exploration Agency
(ISAS/JAXA; $39.16^\circ$N, $141.82^\circ$E) on September 1, 2006 at 06:11 Japan Standard Time (JST).
The flight path is shown in Fig.~\ref{fig:flight_path}.
Figure~\ref{fig:altitude} shows the time variations of the balloon's altitude and atmospheric pressure.
At 08:56, the balloon reached an altitude of $35.0$ km and began level flight.
From 08:56 to 10:15,
the altitude was constant at $35$ km and the atmospheric pressure was $5.4$ hPa ($5.5$ g cm$^{-2}$),
and from 11:20 to 13:00, the altitude was $32$ km and the pressure was $8.5$ hPa ($8.7$ g cm$^{-2}$).
During level flight between 08:56 and 13:00,
the time average of the atmospheric depth was $7.0$ g cm$^{-2}$,
the live time of the observation was $3.0$ hr,
and the cutoff rigidity calculated by QARM was $9.7$ GV.
Between 12:06 and 12:33,
we suspended gamma-ray observation, and the camera was operated in the charged-particle tracking mode,
which was used to check the performance of the tracker with a trigger of any two hits on GSO pixels.
At 12:33, we resumed gamma-ray observation.
Then, we turned off the system power at 12:59.
Finally, we cut the gondola from the balloon at 13:20.
The gondola landed in the sea at 13:45,
and we successfully recovered it at 14:32.
During the whole flight,
no serious problem occurred with the balloon system or the detector.
\begin{figure}
\plotone{fig5.eps}
\caption{
Flight path of SMILE-I.
The balloon was launched from the Sanriku Balloon Center (SBC) of ISAS/JAXA in Japan (39.16N, 141.82E),
and it landed in the sea approximately $30$ km from the SBC.
\label{fig:flight_path}
}
\end{figure}
\begin{figure}
\plotone{fig6.eps}
\caption{
Time variation of the balloon's altitude and atmospheric pressure.
\label{fig:altitude}
}
\end{figure}
\section{Results \& Discussion}
\subsection{Event Reconstruction}
In this flight, $2.2 \times 10^5$ events were acquired during the $6.8$ hr after the launch.
If an incident gamma ray had been scattered in the TPC and then hit more than two pixels in the GSO:Ce absorber,
we could not reconstruct the gamma-ray event because the sequence of interactions in the absorber was unknown.
We, therefore, selected events with a single hit in the absorber,
and $1.1 \times 10^5$ events remained.
Figure~\ref{fig:e_tr} shows the energy deposit in the TPC and the track length of those events.
If a Compton-recoil electron stops in the TPC,
the event is plotted around the dashed line in this figure, which represents the relationship obtained from the Geant4 simulation.
On the other hand,
if the electron escaped from the TPC,
the event is located around the solid line, which represents the relationship calculated from energy deposit per unit length for minimum ionizing particles.
This is because with increasing energy of the recoil electron its energy deposit per unit length of the electron escaping
from the TPC approaches that of minimum ionizing particles.
Because the energy of a Compton-recoil electron is necessary for the gamma-ray reconstruction,
$6.5 \times 10^3$ events remained after we selected events in the case that the electron stopped in the TPC.
We used the following criteria to select events in which the electron stopped in the TPC:
\begin{eqnarray}
& \left| \frac{L_e}{\rm [cm]} - 1.8\times 10^{-3} \left( \frac{K_e}{\rm [keV]} \right) ^{1.8} \right| \leq 2 \label{eq:e_tr}, \\
& K_e > 15 \ {\rm [keV]} \label{eq:e_lim},
\end{eqnarray}
where $L_e$ is the length of the detected track in the tracker.
We know from our experience with ground-based experiments using the prototype trackers
that an electron deposits all its energy in the TPC if it satisfies equation~(\ref{eq:e_tr}).
The energy spectrum of the TPC has a line component at $8$ keV
which corresponds to the characteristic CuK$\alpha$ X-ray line from the copper electrode of the TPC.
Because these events are not Compton events, we used the threshold given in equation~(\ref{eq:e_lim}) to reject them.
Thus, we selected Compton events in which the electron stopped in the TPC in order to be able to reconstruct the incident gamma rays.
Finally, we selected events whose Compton scattering occurred in the fiducial volume of $9 \times 9 \times 13$ cm$^3$
by using the angle $\alpha$ and equation~(\ref{eq:alpha_cut}) with $\Delta_\alpha$ of
\begin{equation}
\Delta_\alpha = 0.075 \times \frac{(E_\gamma + K_e)^2}{K_e \left| E_g - m_ec^2 \right|} \label{eq:d_alpha},
\end{equation}
$2.1 \times 10^3$ events then remained.
This criteria of the angle $\alpha$ is equivalent to $| 1 - |{\bf r}|^2 | \leq 0.15$,
where $\bf r$ was described by equation~(\ref{eq:kin_rcs})
and $|\bf r|$ is equal to unity under the condition of $\cos \alpha_{geo} = \cos \alpha_{kin}$.
The time variation of the count rate of the reconstructed events is shown in Fig.~\ref{fig:rate_flight},
which has a maximum near the Pfotzer maximum
and is thereafter constant during level flight, with a standard deviation of $20$\%.
During level flight,
we obtained $881$ reconstructed events in the detected energy range of $100$ keV--$1$ MeV,
$420$ events of which were downward at zenith angle less than $60$ degrees.
The energy spectra of both the downward events and the events in all directions are shown in Fig.~\ref{fig:spec_level_flight}.
\begin{figure}
\plotone{fig7.eps}
\caption{
Track length and deposited energy in the TPC for events that occurred during level flight (scatter plot).
The solid and dashed lines represent the relationship calculated from energy deposit per unit length of minimum ionizing particles
and the relationship simulated using Geant4 for electrons stopped in the TPC, respectively.
The hatched area represents the event selection described by equations~(\ref{eq:e_tr}) and (\ref{eq:e_lim}).
\label{fig:e_tr}
}
\end{figure}
\begin{figure}
\plotone{fig8.eps}
\caption{
Time variation of the rate of reconstruction of events.
\label{fig:rate_flight}
}
\end{figure}
\begin{figure}
\plotone{fig9.eps}
\caption{
Energy spectra of the events that were reconstructed as gamma rays during level flight in all directions (circles),
and at zenith angles below $60$ degrees (squares).
\label{fig:spec_level_flight}
}
\end{figure}
\subsection{Instrumental Background Simulation}
In a balloon-borne experiment,
particles incident on a detector are not only gamma rays but also protons, electrons, neutrons, and others.
These particles and the gamma rays induced by interactions between these particles
and the pressurized vessel or materials surrounding the detector could form a part of the background.
In particular,
a neutron causes elastic scattering,
so that when a neutron interacts with both the scatterer and the absorber of a Compton camera,
the neutron event may be incorrectly reconstructed as a Compton-scattered gamma ray in Compton-type detectors.
The electron tracker of an ETCC can distinguish between electrons stopped in the TPC and other charged particles
because of the difference in the track length as a function of the kinetic energy, as shown in Fig.~\ref{fig:e_tr}.
The purpose of the SMILE-I experiment was the measurement of diffuse cosmic and atmospheric gamma rays.
Therefore, atmospheric gamma rays produced in the atmosphere by cosmic-ray interactions are considered not as background but as signals in this experiment.
\begin{figure}
\plotone{fig10.eps}
\caption{
Mass model for the instrumental background simulation.
The model of the electron-tracking Compton camera (ETCC) is the same as in Fig.~\ref{fig:mass_model_etcc}.
\label{fig:mass_model_vessel}
}
\end{figure}
To estimate background events,
we calculated the background radiation at the balloon's altitude by using a Geant4 simulation and a mass model shown in Fig.~\ref{fig:mass_model_vessel}.
The balloon gondola was not included in the mass model,
because most of the parts of the balloon gondola were expanded polystyrene and the effect on secondary particle production was negligible.
Due to the low density of the electronics boxes and the large distance between the ETCC and the battery,
the effect of the instrumental background can be neglected.
Therefore, this mass model does not include a gondola, electronics box, and battery.
The Geant4 hadronic processes were elastic and inelastic scattering for the low- and high-energy models,
and absorption for pions, kaons, protons, anti-protons, and anti-neutrons;
elastic and inelastic scattering, and capture with a high-precision model below $19$ MeV for neutrons;
elastic and inelastic scattering with a low-energy model for deutrons, tritons, and alpha particles.
In addition, {\it G4Decay} and {\it G4RadioactiveDecay} processes were included for prompt processes and delayed photons from radioactive decays.
We did not include the chance coincidences between triggers from different components of the radiation background,
since their contribution is expected to be small by the background rejection method described in section 2.
As described in section 1,
there are several radiation models for the fluxes of incident particles at balloon altitudes.
Mizuno's model is a radiation model based on the previous observations
of protons and alpha particles above $100$ MeV, electrons and positrons above $100$ MeV,
gamma rays between $1$ MeV and $100$ GeV, and muons between $300$ MeV and $20$ GeV.
However, below these energy ranges,
the flux of each of these particles is simply assumed to be a power-law spectrum with an index of $-1$,
and the neutron flux is not described at all in this model.
In contrast, the QARM model, based on previous observations and some simulations,
has fluxes of neutrons in addition to protons, electrons, gamma rays, pions, and muons in wider energy ranges between $100$ keV and $100$ GeV
and includes the effect of solar modulation on the spectral shape and normalization.
The parameters of the QARM model are date,
geographic position, altitude/residual pressure, magnetic field disturbance index, and the primary-particle spectrum.
Because the model of low-energy charged particles and neutrons was important in sub-MeV gamma-ray observations,
we adopted the QARM spectra shown in Fig.~\ref{fig:back_inc},
assuming a primary-particle spectrum of galactic cosmic rays \citep{Lei_2004}
under a magnetic field disturbance index $K_p$ \citep{Mayaud_1980} of $3$
and an atmospheric depth of $7$ g cm$^{-2}$.
The zenith angle distribution of QARM for public use is divided in only two regions: the zenith angle range of $0$--$90$ degrees, and $90$--$180$ degrees.
Figure~\ref{fig:back_inc} represents the downward fluxes in the zenith angle range of $0$--$90$ degrees and the upward fluxes in the zenith angle range of $90$--$180$ degrees.
We estimated an uncertainty in the particle fluxes of QARM to be $20$\%
by varying the atmospheric depth between $5.3$ g cm$^{-2}$ and $8.5$ g cm$^{-2}$
and the magnetic field disturbance index between $2$ and $4$.
\begin{figure}
\plotone{fig11.eps}
\caption{
Energy spectra of protons (solid lines), electrons (dashed lines), and neutrons (dotted lines) for a background simulation
at an atmospheric depth of $7.0$ g cm$^{-2}$.
Red shows the downward flux in the zenith angle range of 0-90 degrees and blue shows the upward flux in the zenith angle range of 90-180 degrees.
All fluxes were calculated by QARM.
\label{fig:back_inc}
}
\end{figure}
For the simulation of the instrumental background coming from any direction,
we assume the incident direction of the primary particles to be downward (zenith angle = $0$ degrees) or upward (zenith angle = $180$ degrees),
because we do not have any observational data of the horizontal fluxes or the zenith angle distributions of protons, neutrons, and electrons.
For the incident fluxes,
we adopted the QARM flux in the zenith angle range of $0$--$90$ degrees (Fig.~\ref{fig:back_inc}) as the downward primary particle flux,
and the QARM flux in the zenith angle range of $90$--$180$ degrees (Fig.~\ref{fig:back_inc}) as the upward primary particle flux.
The simulated spectra of instrumental background gamma rays produced in the materials surrounding the detector are shown in Fig.~\ref{fig:bg_photon}.
The spectra shown in Fig.~\ref{fig:bg_photon} are the flux averaged over the solid angle in the zenith angle range between $0$ and $90$ degrees.
Gamma rays induced by electrons dominate in the energy range of SMILE-I,
and neutron-induced gamma rays contribute slightly at several MeV.
In addition, there is a line component of electron--positron annihilation at $511$ keV.
In order to estimate the contribution of secondary gamma rays, neutrons, and charged particles incoming to the gaseous tracker from all directions,
we simulated their fluxes (Fig.~\ref{fig:bg_particles}).
Throughout this paper, secondary gamma rays refer to photons due to the instrumental background only.
Secondary gamma rays contribute the majority of the background,
with the rest coming from neutrons, electrons, positrons, protons, and charged pions.
The rejection inefficiencies for background particles
after Compton reconstruction under the constraints of equations~(\ref{eq:e_tr}), (\ref{eq:e_lim}), and (\ref{eq:d_alpha}) were
estimated by the simulation to be $1.8\pm0.3 \times 10^{-5}$ for neutrons,
$3.4\pm0.2 \times 10^{-4}$ for electrons, and $2.5\pm0.2 \times 10^{-4}$ for protons.
Therefore,
the simulation predicted that the background event rates of gamma rays, neutrons, and charged particles during level flight would be
$7.4\pm2.6 \times 10^{-3}$ s$^{-1}$, $6.7\pm2.6 \times 10^{-4}$ s$^{-1}$, and $8.7\pm2.3 \times 10^{-5}$ s$^{-1}$, respectively.
The uncertainties are caused by the uncertainty in the background model.
Because the rate of reconstructed events during the level flight of SMILE-I was $3.9 \times 10^{-2}$ s$^{-1}$,
the contributions of background gamma rays, neutrons, and charged particles to the obtained events were
$19$\%, $1.7$\%, and $0.23$\%, respectively.
Thus, we estimated that $98.0 \pm 0.7$\% of the reconstructed events were gamma-ray events,
and the contribution of the other particles was $2.0 \pm 0.7$\%,
thanks to the powerful background rejection of the ETCC.
Figure~\ref{fig:bg_depth} shows the source function of instrumental secondary gamma rays incoming to the detector
from the zenith direction as a function of atmospheric depth.
These spectra do not include any event selection described in section 4.1.
In later analysis,
we use this flux of gamma rays as the background.
\begin{figure}
\plotone{fig12.eps}
\caption{
Simulated source function of the secondary photons produced in the instrument
by protons (dotted line), electrons (dashed line), and neutrons (long dashed line)
at zenith angles between $0$ and $90$ degrees,
spectra of which are shown in Fig.~\ref{fig:back_inc},
at an atmospheric depth of $7.0$ g cm$^{-2}$.
The solid line shows the total flux.
\label{fig:bg_photon}
}
\end{figure}
\begin{figure}
\plotone{fig13.eps}
\caption{
Simulated differential fluxes of particles incident on the gaseous tracker from all directions
for secondary gamma rays (open circles), neutrons (crosses), electrons (squares),
positrons (asterisks), protons (filled circles), $\pi^+$ (stars), and $\pi^-$ (triangles),
at an atmospheric depth of $7.0$ g cm$^{-2}$.
These spectra include both primary (shown in Fig.~\ref{fig:back_inc}) and secondary particles.
\label{fig:bg_particles}
}
\end{figure}
\begin{figure}
\plotone{fig14.eps}
\caption{
Differential flux of instrumental background photons
produced by protons (dotted line), electrons (dashed line), and neutrons (long dashed line) in the $125$--$1250$ keV region,
as a function of atmospheric depth.
These fluxes represent the background source function averaged in the zenith angle between 0 degree and 90 degrees at the sensitive area of the ETCC.
The solid line shows the total flux.
\label{fig:bg_depth}
}
\end{figure}
\subsection{Gamma-ray Scattering in the Atmosphere}
In observing diffuse cosmic gamma rays at a few hundred keV,
there is a contribution of cosmic gamma rays
that are scattered into the aperture of the detector
after one or more Compton scatterings in the overlying atmosphere.
The cosmic gamma-ray flux $f_c (z, \theta)$ at the atmospheric depth of $z$ and the zenith angle of $\theta$ can be separated into two components,
\begin{equation}
f_c (z, \theta) = f_n (z, \theta) + f_s (z, \theta) .
\label{eq:cosmic}
\end{equation}
$f_n (z, \theta)$ is the flux of cosmic gamma rays
that enter the aperture of the detector without any interaction in the overlying atmosphere.
In the zenith direction ($\theta = 0$),
the depth dependence of $f_n (z, \theta)$ is given by
\begin{equation}
f_n (z, 0) = f_n(0, 0) \exp(- \tau_{tot} z) = \eta \exp(- \tau_{tot} z) ,
\label{eq:cs_att}
\end{equation}
where $\eta$ is the flux of cosmic gamma rays at the top of the atmosphere,
and $\tau_{tot}$ is the cross section of the total attenuation per unit mass.
On the other hand,
$f_s (z, \theta)$ is the flux of cosmic gamma rays
scattered into the aperture of the detector
after one or more Compton scatterings in the overlying atmosphere.
To obtain the ratio of the scattered component to the total diffuse cosmic gamma-ray flux,
we simulated the transport of gamma rays in the atmosphere with Geant4.
We defined $80$ atmospheric layers from sea level to an altitude of $80$ km,
where each layer was a $1$-km thick sphere,
filled with air with the density appropriate for each altitude.
Initial gamma rays were generated at random positions on the surface of a sphere of $650$ km radius and given random initial directions,
and we traced their histories.
Then, the ratio of the scattered component was calculated
as a function of energy, atmospheric depth, and zenith angle.
Figures~\ref{fig:correction_factor} and \ref{fig:c_factor_various_depth} show
the scattered component ratio $\lambda (z, \theta)$ as a function of the detected energy,
\begin{equation}
\lambda (z, \theta) = \frac{f_s (z, \theta)}{f_n (z, \theta) + f_s (z, \theta)} ,
\label{eq:correction}
\end{equation}
assuming a photon index of $2.0$ in the primary spectrum.
\begin{figure}
\plotone{fig15.eps}
\caption{
Ratio of the scattered component $\lambda$ of diffuse cosmic gamma rays to the total (scattered plus nonscattered) gamma rays
as a function of the energy
at zenith angles between $0$ and $20$ degrees
at an atmospheric depth of $8.0$ g cm$^{-2}$.
The solid and dotted lines show our simulation results
with and without Rayleigh scattering, respectively.
The long-dashed line, dashed line, and stars show
the simulation by \citet{Horstman_1971},
the calculation by \citet{Makino_1970},
and the simulation by \citet{Schonfelder_1977}, respectively.
Each model is at an atmospheric depth of $8.0$ g cm$^{-2}$.
\label{fig:correction_factor}
}
\end{figure}
\begin{figure}
\plotone{fig16.eps}
\caption{
Our simulation results with Rayleigh scattering of the ratio of the scattered component
$\lambda$ of diffuse cosmic gamma rays as a function of the energy
at zenith angles between $0$ and $20$ degrees
at atmospheric depths of $52.6$, $11.5$, $2.78$, and $0.78$ g cm$^{-2}$ (upper to lower lines).
There are five data sets for each curve: $10$--$100$ keV, $100$ keV--$1$ MeV, $500$ keV--$5$ MeV, $1$--$10$ MeV, and $2.5$--$25$ MeV,
and the points in overlapped energy range are overlaid.
\label{fig:c_factor_various_depth}
}
\end{figure}
There are two breaks in Fig.~\ref{fig:correction_factor}.
The first break is at around $30$ keV, where absorption and scattering dominate below and above the peak energy, respectively.
The energy of the first break is constant at any atmospheric depth, as shown in Fig.~\ref{fig:c_factor_various_depth}.
The second break in Fig.~\ref{fig:correction_factor} appears at approximately $1$ MeV.
The energy of the second break decreases as the atmospheric depth decreases, as shown in Fig.~\ref{fig:c_factor_various_depth},
and the optical depth at the second break is approximately one.
Therefore, for gamma rays with energies above the second break, the probability of scattering is so low that $\lambda$ decreases.
Moreover, Rayleigh scattering contributes a large amount (more than $10$\%) to the scattering component ratio under $50$ keV.
Compared with the previous calculation and simulations of the scattered component ratio function,
our simulation result has a difference of $\leq 10$\%
from the simulation of Horstman at $30$--$100$ keV \citep{Horstman_1971},
$\sim 20$\% from the calculation of Makino at hundreds of keV \citep{Makino_1970},
and $\sim 10$\% from the simulation of Sch\"onfelder above $1$ MeV \citep{Schonfelder_1977}.
Makino's calculation assumed
that the scattering angle dependence of the cross section was negligible,
and only one scattering occurred.
Therefore,
Makino's calculation differs considerably from our result
in the energy range under $100$ keV because of the scattering number limitation,
and it also differs above $100$ keV
because it neglects the scattering angle dependence of the cross section.
The contribution of cosmic gamma rays that are scattered in the atmosphere is not negligible in the sub-MeV/MeV region;
thus, we also used Geant4 to simulate the scattering component ratio as a function of energy (for various power indices),
atmospheric depth, and zenith angle (Fig.~\ref{fig:c_factor_index}--~\ref{fig:c_factor_zenith}).
Figure~\ref{fig:c_factor_index} shows that the ratios for different power indices differ by less than $5$\%
in the range between tens of keV and a few MeV;
thus, the choice of the power index does not strongly affect our results.
\begin{figure}
\plotone{fig17.eps}
\caption{
Scattering component ratio $\lambda$
for the initial spectra with power indices of $2.0$ (solid line) and $2.5$ (dashed line)
at an atmospheric depth of $7.0$ g cm$^{-2}$ and zenith angles between $0$ and $60$ degrees.
\label{fig:c_factor_index}
}
\end{figure}
Figure~\ref{fig:c_factor_depth} shows that the atmospheric depth dependence of $\lambda $ is similar to that of Makino's calculation,
$\case{1}{1-\lambda} \propto \log (1 + \tau_{tot} z)$,
although there is a $10$\%--$20$\% difference in the energy range of $100$--$500$ keV.
Figure~\ref{fig:c_factor_zenith} shows that although the zenith dependence of $\lambda$ below a few MeV is weak
in the region of zenith angles less than $50$ degrees, it changes abruptly above $70$ degrees.
Thus, the zenith angle dependence below a few MeV is not consistent with $1/\cos \theta$,
where $\theta$ is the zenith angle of the incident gamma ray.
For the higher energies, $\lambda$ is similar to $1/\cos \theta$.
\begin{figure}
\plotone{fig18.eps}
\caption{
Ratio of the scattering component $\lambda$ as a function of the atmospheric depth
at zenith angles between $0$ and $5$ degrees,
in the energy ranges $30$--$50$ keV, $100$--$150$ keV, $300$--$500$ keV, $750$--$1000$ keV,
$1000$--$1500$ keV, $3000$--$5000$ keV, and $5000$--$10000$ keV.
\label{fig:c_factor_depth}
}
\end{figure}
\begin{figure}
\plotone{fig19.eps}
\caption{
Ratio of the scattering component $\lambda$ as a function of the zenith angle
in the energy ranges of $30$--$50$ keV, $100$--$150$ keV, $300$--$500$ keV,
$750$--$1000$ keV, $1000$--$1500$ keV, $3000$--$5000$ keV, and $5000$--$10000$ keV.
The dashed line is proportional to $1/\cos \theta$.
\label{fig:c_factor_zenith}
}
\end{figure}
\subsection{Diffuse Cosmic and Atmospheric Gamma-ray Flux}
To study the origin of the reconstructed events,
we divided the obtained data at atmospheric depths between $5.2$ g cm$^{-2}$ and $168$ g cm$^{-2}$ into $10$ segments as Table~\ref{tab:time_segment}.
Then,
we obtained the gamma-ray flux from the zenith direction ($\theta = 0$--$10$ degrees) for each time segment
by analyzing the reconstructed events with a simulated response matrix
for incident energies of 100, 150, 200, 300, 400, 511, 600, 1000, and 1500 keV
and incident zenith angles of 0, 20, 40, 60, 80, 120, 150, and 180 degrees.
In the calculation,
the zenith angle dependences of both diffuse cosmic and atmospheric gamma-ray fluxes must be considered,
because our ETCC had a wide FOV of $3$ sr.
For diffuse cosmic gamma rays, we simulated the zenith angle dependence as described in section~4.3.
For the zenith angle distribution of combined cosmic and atmospheric gamma rays,
only three models exist and the conditions under which they apply are limited, as follows.
1) In Ling's model, the source function, which is the production rate of atmospheric gamma rays at the generated position,
is defined on the basis of observed data but has no dependence on the zenith angle.
In transport, only absorption is included, and scattering is not considered.
Ling's model also describes the flux of diffuse cosmic gamma rays between $300$ keV and $10$ MeV at atmospheric depths smaller than $500$ g cm$^{-2}$,
which is comparable to the results of several experiments,
although the upward gamma-ray flux differs by a factor of three from the observed data \citep{Schonfelder_1977}.
2) Graser's model calculates atmospheric gamma rays based on bremsstrahlung using some models of the electron flux.
This model ignores scattering in transport and $\pi^0$-decay gamma rays.
Thus, this model is effective in the energy range of $1$--$30$ MeV.
3) Costa's model is an empirical model for the sum of cosmic and atmospheric gamma rays
based on observations by \citet{Schonfelder_1977, Schonfelder_1980} and \citet{Ryan_1977}.
This model fits the experimental data well but is applicable only at atmospheric depths under $10$ g cm$^{-2}$.
Given the limitations of each of these models,
it is difficult to choose which one best describes the zenith angle distribution of combined cosmic and atmospheric gamma rays.
As described in section~4.3,
the contribution of the scattering of cosmic gamma rays in the atmosphere is not negligible within the energy range of SMILE-I.
Similarly, the scattering of atmospheric gamma rays in the atmosphere is not negligible.
However, Graser's model ignores scattering in the atmosphere.
Ling's model also does not include any scattering in transport.
However, the difference from the observed zenith-angle dependence is less than that of Graser's model,
because the source function of Ling's model was fitted to the sum fluxes of cosmic and atmospheric gamma rays obtained from previous observations.
Moreover, Ling's model covers the range of atmospheric depths corresponding to all altitudes of the SMILE-I flight.
We, therefore, adopted Ling's model including the cosmic component.
However, the Ling's model describes the flux only above $300$ keV.
Thus, we extended Ling's model,
which was extrapolated to less than $300$ keV with a single power-law spectrum fitted between $300$ keV and $5$ MeV.
Below an atmospheric depth of $10$ g cm$^{-2}$,
the difference between the extended Ling's model and Costa's model in the energy range below $300$ keV is less than $\pm 5$\% at all zenith angles.
However, we could not find a report on any observational data below $300$ keV
that could be compared with the extended Ling's model above the atmospheric depth of $10$ g cm$^{-2}$.
Therefore, although we adopted an extended Ling's model,
we could not estimate its uncertainty.
\begin{deluxetable*}{cccccc}
\tablecolumns{6}
\tablewidth{0pc}
\tablecaption{Obtained gamma-ray fluxes in the zenith angle range of $0$--$10$ degrees for each time segment. \label{tab:time_segment}}
\tablehead{
\colhead{Time} &\colhead{Atmospheric depth} &\multicolumn{4}{c}{Flux [$10^{-4}$ photons s$^{-1}$ cm$^{-2}$ sr$^{-1}$ keV$^{-1}$]} \\
\colhead{[JST]} &\colhead{[g cm$^{-2}$]} &\colhead{125--250 keV} &\colhead{250--550 keV} &\colhead{550--1250 keV} &\colhead{125--1250 keV}
}
\startdata
07:20--07:47 &145 (121--168) &$8.8\pm1.2$ &$5.5\pm0.6$ &$1.6\pm0.2$ &$3.4\pm0.3$\\
07:47--08:00 &95 (70--121) &$18\pm10$ &$9.0\pm1.2$ &$2.6\pm0.4$ &$6.0\pm1.2$\\
08:00--08:12 &55 (40--70) &$12\pm 2$ &$7.4\pm1.0$ &$2.1\pm0.4$ &$4.6\pm0.5$\\
08:12--08:24 &32 (23--40) &$7.1\pm1.1$ &$4.5\pm0.5$ &$1.3\pm0.2$ &$2.8\pm0.3$\\
08:24--08:56 &12 (5.3--24) &$5.1\pm0.7$ &$3.2\pm0.4$ &$0.92\pm0.13$ &$2.0\pm0.2$\\
08:56--10:08 &5.4 (5.2--5.6) &$3.0\pm0.4$ &$1.9\pm0.2$ &$0.53\pm0.07$ &$1.1\pm0.1$\\
10:08--11:00 &6.8 (5.4--7.9) &$3.9\pm0.5$ &$2.4\pm0.3$ &$0.69\pm0.09$ &$1.5\pm0.1$\\
11:00--11:30 &8.2 (7.7--8.6) &$3.7\pm0.5$ &$2.3\pm0.3$ &$0.65\pm0.10$ &$1.4\pm0.1$\\
11:30--12:06 &8.5 (8.3--8.7) &$3.8\pm0.6$ &$2.4\pm0.3$ &$0.67\pm0.10$ &$1.5\pm0.1$\\
12:33--12:59 &8.3 (8.0--8.7) &$3.2\pm2.3$ &$2.0\pm1.2$ &$0.57\pm0.33$ &$1.2\pm0.5$
\enddata
\end{deluxetable*}
We used this extended Ling's model of the cosmic and atmospheric gamma rays along with the simulated response matrices
to unfold the observed events with reconstructed directions in the zenith angle below $60$ degrees,
resulting in the gamma-ray fluxes listed in Table~\ref{tab:time_segment}.
In this analysis, we assumed
that the zenith angle dependence of the sum of diffuse cosmic, atmospheric, and instrumental gamma rays was the same as the extended Ling's model.
Using this data,
the growth curve, which is the dependence of the gamma-ray flux on atmospheric depth, in the energy range of $125$--$1250$ keV was obtained,
as shown in Fig.~\ref{fig:growth_curve}.
For comparison with the extended Ling's model,
the atmospheric gamma-ray flux was considered to have a dependence on $R_{cut}^{-1.13}$,
where $R_{cut}$ is the cutoff rigidity \citep{Thompson_1981}.
Because the cutoff rigidities of the SMILE-I flight and Ling's model are $9.7$ GV and $4.5$ GV, respectively,
the sum of the atmospheric gamma-ray flux of the extended Ling's model multiplied by $(9.7/4.5)^{-1.13}$
and the diffuse cosmic gamma-ray flux of the extended Ling's model between $125$ keV and $1250$ keV
are also shown in Fig.~\ref{fig:growth_curve}.
The SMILE-I result was consistent with the sum of the extended Ling's atmospheric and cosmic model
and the background gamma-ray flux represented by the solid line in Fig.~\ref{fig:bg_depth}.
At atmospheric depths from $5$ to $100$ g cm$^{-2}$,
the variability of the ratio of the background particle flux to the cosmic and atmospheric gamma-ray flux is slight ($< \pm 5$\%).
Thus, the ratio of the non-Compton-scattering events of neutrons or charged particles to the reconstructed events is
approximately equal to that of the level flight, $2$\%, which was estimated in section 4.2.
In other words, $98$\% of the reconstructed events were induced by cosmic, atmospheric, and instrumental background gamma rays.
Therefore, we calculated the fluxes of cosmic and atmospheric gamma rays by using the fluxes obtained from the SMILE-I observations
assuming that the reconstruction of non Compton-scattering events was negligible.
\begin{figure}
\plotone{fig20.eps}
\caption{
Reconstructed gamma-ray flux in the zenith angle range of $0$ to $10$ degrees
in the energy range $125$--$1250$ keV as a function of atmospheric depth.
The circles represent data taken with the balloon ascending.
The dashed line shows the sum of 1) the simulated background (solid line in Fig.~\ref{fig:bg_depth}),
2) the diffuse cosmic gamma-ray flux of our extended Ling's model between $125$ and $1250$ keV,
and 3) the atmospheric gamma-ray flux of the same model.
The atmospheric component of the extended Ling's model is scaled by the cutoff rigidity.
\label{fig:growth_curve}
}
\end{figure}
\begin{figure}
\plotone{fig21a.eps}
\plotone{fig21b.eps}
\plotone{fig21c.eps}
\caption{
Separation of atmospheric and cosmic components of gamma rays
by fitting the growth curves in the energy ranges (a) $125$--$250$ keV,
(b) $250$--$550$ keV, and (c) $550$--$1250$ keV,
in the zenith angle range of $0$--$10$ degrees.
The long-dashed, dotted, dot-dashed, and dashed lines represent
the contributions from cosmic gamma rays, atmospheric gamma rays, instrumental background components,
and the electron--positron annihilation line, respectively.
The solid line represents their sum.
\label{fig:growth_fit}
}
\end{figure}
\begin{deluxetable*}{ccc}
\tablecolumns{3}
\tablewidth{0pc}
\tablecaption{
Obtained diffuse cosmic and atmospheric gamma-ray fluxes in the zenith direction ($0$--$10$ degrees).
\label{tab:flux_eng}
}
\tablehead{
\colhead{Energy} &\colhead{Cosmic gamma rays\tablenotemark{$\dagger$}} &\colhead{Atmospheric gamma rays\tablenotemark{$\dagger$}} \\
\colhead{[keV]} &\colhead{[ph s$^{-1}$ cm$^{-2}$ sr$^{-1}$ keV$^{-1}$]}
&\colhead{[ph s$^{-1}$ cm$^{-2}$ sr$^{-1}$ keV$^{-1}$ (g cm$^{-2}$)$^{-1}$]}
}
\startdata
125--250 &$2.1 \pm 0.7 ^{+ 1.0}_{- 0.8} \times 10^{-4}$ &$2.0 \pm 0.6 ^{+ 0.5}_{- 0.9} \times 10^{-5}$\\
250--550 &$8.3 \pm 3.4 ^{+ 5.5}_{- 4.3} \times 10^{-5}$ &$7.8 \pm 2.9 ^{+ 3.0}_{- 5.5} \times 10^{-6}$\\
550--1250 &$3.3 \pm 1.2 ^{+ 1.6}_{- 1.2} \times 10^{-5}$ &$3.3 \pm 1.0 ^{+ 0.9}_{- 1.6} \times 10^{-6}$
\enddata
\tablenotetext{$\dagger$}{The first uncertainty term is statistical and the second is systematic.}
\end{deluxetable*}
To obtain the fluxes of diffuse cosmic gamma rays and atmospheric gamma rays,
we used the difference in the atmospheric depth dependence of each component.
For the cosmic gamma-ray flux at an atmospheric depth of $z$,
we assumed
\begin{equation}
f_c (z) = \frac{\eta}{1 - \lambda (z, 0)} \exp (- \tau_{tot} z) ,
\end{equation}
with equations (\ref{eq:cosmic}), (\ref{eq:cs_att}), and (\ref{eq:correction}) at $\theta = 0$.
On the other hand,
the atmospheric gamma-ray flux $f_a (z)$ is nearly proportional to the atmospheric depth \citep{Schonfelder_1977},
\begin{equation}
f_a (z) = \xi z .
\end{equation}
Then, we fitted the observed growth curve to
\begin{equation}
f(z) = \xi^\prime z + \frac{\eta^\prime}{1 - \lambda (z, 0)} \exp ( -\tau_{tot} z) + f_b (z) + F_{511} (z) \delta (E_{0} - 511 \ \rm{[keV]}) ,
\label{eq:growth}
\end{equation}
where $f_b$ is the flux of background photons estimated by the Geant4 simulation (Fig.~\ref{fig:bg_depth}),
$F_{511}$ is the flux of the atmospheric annihilation line described by equation~(1) in \citet{Ling_1977},
and $\xi^\prime$ and $\eta^\prime$ are the attenuated fluxes through the pressurized vessel.
The attenuated fluxes are given as $\xi^\prime = \xi \exp ( -\tau_{abs} d l)$ and $\eta^\prime = \eta \exp ( -\tau_{abs} d l)$,
where $\tau_{abs}$ is the cross section of photoelectric absorption in aluminum,
$d = 2.7$ g cm$^{-3}$ and $l = 3$ mm are the density and the thickness of the pressurized vessel, respectively.
The growth curves with the fitting functions are shown in Fig.~\ref{fig:growth_fit},
and these functions serve to isolate individual contributions to the total flux
so that separate fluxes for diffuse cosmic gamma rays and atmospheric gamma rays can be obtained.
However, at larger atmospheric depths than approximately $100$ g cm$^{-2}$,
we have a problem because the gamma-ray flux is not proportional to the atmospheric depth, as shown by Ling's model in Fig.~\ref{fig:growth_curve}.
Therefore, the data at the highest atmospheric depth was not used in the fitting.
The obtained fluxes of cosmic and atmospheric gamma rays are listed in Table~\ref{tab:flux_eng}.
The systematic uncertainties in the cosmic and atmospheric gamma-ray fluxes are caused by the following:
1) The uncertainty of $15$\% in the detection efficiency (section 2).
2) The difference of $5$\% in the relative ratio of the zenith angle distributions between the extended Ling's model and Costa's model (section 4.4).
3) The uncertainty of $20$\% in the background gamma-ray flux obtained by varying the parameters in QARM (section 4.2).
4) The uncertainty of $5$\% in the correction factor of scattering in the atmosphere (section 4.3).
Figures~\ref{fig:cosmic_flux} and \ref{fig:atmos_flux} show
the differential flux of diffuse cosmic gamma rays and atmospheric gamma rays, respectively,
with results of previous observations for comparison.
In Fig.~\ref{fig:atmos_flux}, fluxes from observations for which the FOV was not described in the references were not plotted.
In the energy range of $125$--$250$ keV, the flux of atmospheric gamma rays obtained by SMILE-I
had a difference of $3$ sigma from the observation of \citet{Peterson_1972}.
However, other results for the diffuse cosmic and atmospheric fluxes obtained by SMILE-I were consistent
with previous observations and the Ling and QARM models of atmospheric gamma rays within $2$ sigma.
\begin{figure*}
\plotone{fig22.eps}
\caption{
Differential fluxes of diffuse cosmic gamma rays as a function of energy.
Open squares denote our results, and the uncertainty bars represent the $1\sigma$ combined statistical and systematic uncertainties.
References to previous observations are provided in the text.
\label{fig:cosmic_flux}
}
\end{figure*}
\begin{figure*}
\plotone{fig23.eps}
\caption{
Differential fluxes of downward atmospheric gamma rays in the zenith direction ($0$--$10$ degrees), multiplied by $(R_{cut}/9.7 \ \rm{GV})^{1.13}$
as a function of energy.
Open squares denote our results,
and the uncertainty bars represent the $1\sigma$ combined statistical and systematic uncertainties.
The dashed line and stars represent Ling's model and QARM, respectively.
References to previous observations are provided in the text.
\label{fig:atmos_flux}
}
\end{figure*}
\subsection{Sensitivity and Future Prospects}
Using the SMILE-I result, we estimated the continuum detection sensitivity of this camera.
Diffuse cosmic and atmospheric gamma rays, and those induced by charged particles and neutrons,
which were observed in the SMILE-I experiment, would form the background when we observe a celestial point source.
The minimal detectable flux $F_{min}$ at the significance of $3 \sigma$ is described as
\begin{equation}
F_{min} = 3 \ \sqrt{\frac{f \ \Delta E \ \Delta \Omega}{A_{eff} \ T_{obs}}} ,
\end{equation}
where $f$ is the energy spectrum of background radiation in units of photons s$^{-1}$ cm$^{-2}$ sr$^{-1}$ MeV$^{-1}$,
and $A_{eff}$, $T_{obs}$, $\Delta E$, and $\Delta\Omega$ are
the effective area, effective observation time, energy resolution, and angular resolution, respectively \citep{Schonfelder_2001}.
From the results of the SMILE-I flight,
the sensitivity of SMILE-I is expected to reach $4.1 \times 10^{-3}$ MeV s$^{-1}$ cm$^{-2}$ at $300$ keV
and $1.2 \times 10^{-2}$ MeV s$^{-1}$ cm$^{-2}$ at $600$ keV, assuming $\Delta E = E$ and $T_{obs} = 10^6$ s (Fig.~\ref{fig:sensitivity}).
For the next flight (hereafter SMILE-II) to observe the Crab Nebula from balloon altitudes,
the ETCC needs upgrading in both the angular resolution and effective area.
As described in equation~(\ref{eq:phi}),
the ARM, which is the accuracy of the scattering angle, depends mostly on the energy resolution of the absorber.
Therefore, improving the energy resolution of the scintillation camera
would give the ETCC better angular resolution.
A good ARM resolution of $4$ degrees (FWHM) at $662$ keV was realized
using a recently developed scintillator, LaBr$_3$:Ce,
with an excellent energy resolution of $3.0$\% and $5.8$\% at $662$ keV
for a monolithic crystal and an array of pixels, respectively \citep{Kurosawa_2009, Kurosawa_2010}.
From this result, the expected ARM resolution would be $1.7$ degrees and $1.3$ degrees (FWHM) at $2$ MeV and above $5$ MeV, respectively.
As for the effective area,
we have developed a larger ETCC based on a TPC with a volume of $30 \times 30 \times 30$ cm$^3$ \citep{Miuchi_2007},
and its gamma-ray imaging has been successful \citep{Ueno_2008}.
With these improvements and the assumption of the instrumental background scaled by the geometrical area of SMILE-I,
it is expected that the ETCC for SMILE-II on a mission to detect the Crab Nebula during a level flight of $10^4$ s
will have approximately ten times higher sensitivity than that of SMILE-I.
We have additional ideas for improving the future ETCC.
First, we are considering inserting an electron absorber between the electron tracker and the scintillators.
In the MeV region,
because the energy of Compton-recoil electrons is as high as an MeV,
the Compton-recoil electrons partly escape from the effective volume of the TPC.
If an absorber that measures the energy of the escaping Compton-recoil electrons is placed between the TPC and scintillators,
the sensitivity in the MeV region can be improved.
Inserting the electron absorber has the added benefit of reducing the coincidence-timing window,
because we can take the coincidence between the electron absorber and scintillators.
Second, we will adopt CF$_4$ gas as the electron tracker.
Compared with Xe gas, the electron diffusion during electron drift and the multiple scattering effect of the recoil electron are both smaller.
Thus, we will be able to trace the electron tracks more accurately.
In addition, the position resolution of the Compton points and the accuracy of the recoil direction should be improved by updating the tracking algorithm.
Last, we will develop the reconstruction algorithm on the basis of a reflexive calculation, such as the likelihood or maximum entropy method.
If we realize an ETCC with a high ARM due to the use of LaBr$_3$ PSAs with a pixel size of $6 \times 6 \times 40$ mm$^3$,
equipped with a TPC with a larger volume of $50 \times 50 \times 50$ cm$^3$ filled with CF$_4$ gas at a higher pressure of $2$ atm,
the sensitivity will become ten times better than that of COMPTEL in the sub-MeV region, as shown in Fig.~\ref{fig:sensitivity}.
Moreover,
when a photon with energy above $1.02$ MeV causes electron--positron pair creation in the TPC,
we can obtain the incident photon's momentum by measuring the momenta of both the electron and positron with the TPC \citep{Orito_2003}.
Therefore, the camera will have better sensitivity above a few MeV than that shown in Fig.~\ref{fig:sensitivity}.
In the near future,
we will experiment on gamma-ray reconstruction by electron--positron pair creation by using an ETCC and verify the result with a simulation,
and also estimate the sensitivity of an ETCC onboard a satellite
where the radiation environment is much different from that at balloon altitudes,
by a performing the full-scale Monte Carlo simulation.
\begin{figure}
\plotone{fig24.eps}
\caption{
Continuum detection sensitivity as a function of energy with an observation time of $10^6$ s at a significance of $3 \sigma$.
For comparison with previous observations, the sensitivities of SMILE were overlaid on Fig.~1 of \citet{Schonfelder_2004}.
The solid and short-dashed lines represent
the sensitivity of an ETCC with a volume of $50 \times 50 \times 50$ cm$^3$ aboard a long-duration balloon (LDB) through Compton scattering,
without and with an absorber of Compton-recoil electrons, respectively.
\label{fig:sensitivity}
}
\end{figure}
\section{Conclusion}
We are developing an ETCC
as an MeV gamma-ray telescope to conduct an all-sky survey in the next generation
with sensitivity that is one order of magnitude higher than that of COMPTEL.
To prepare for future observations onboard a satellite,
we performed a balloon experiment, SMILE.
As the first step of SMILE (SMILE-I),
we launched an ETCC
consisting of a gaseous TPC with a volume of $10 \times 10 \times 14$ cm$^3$ and GSO:Ce scintillation cameras onboard a balloon in 2006
for the observation of diffuse cosmic gamma rays and atmospheric gamma rays.
In this flight,
we successfully detected $2.1 \times 10^3$ reconstructed events from all directions,
$420$ of which were detected in an FOV of $3$ sr during the live time of $3.0$ hr of level flight.
To evaluate the instrumental background,
we estimated particles incident on the detector with simulations using QARM and Geant4;
the reconstructed events included $1.7$\% neutron events, $0.23$\% charged-particle events, and $98$\% gamma-ray events.
Because the contribution of cosmic gamma rays that are scattered in the atmosphere is not negligible in the sub-MeV/MeV region,
we also use Geant4 to simulate the transport in the atmosphere as a function of energy, zenith angle, and atmospheric depth.
Using these simulations,
we analyzed the growth curve in the $125$--$1250$ keV range
and obtained the fluxes of diffuse cosmic gamma rays and atmospheric gamma rays.
Our results are consistent with those of the previous experiments.
The flight of SMILE-I demonstrated
that our ETCC can perform gamma-ray selection and powerful background rejection,
and thus, may improve MeV gamma-ray astronomy by having better continuum sensitivity.
As the next step,
we plan to observe the Crab Nebula by using an ETCC with a $30 \times 30 \times 30$ cm$^3$ TPC to demonstrate gamma-ray imaging.
Subsequently, we plan to observe several celestial sources with a long-duration balloon flight,
and finally aim for an all-sky survey at energies between $150$ keV and $20$ MeV
with sensitivity up to an order of magnitude better than the current and previous missions,
using an ETCC with ARM resolution of $2$ degrees (FWHM) in the MeV region
and a detection volume of $50 \times 50 \times 50$ cm$^3$ onboard a long-duration balloon or satellite.
\acknowledgments
The authors thank the anonymous referee for carefully reading the manuscript
and providing helpful comments to improve it.
We also thank the staff of the Scientific Balloon Laboratory, ISAS/JAXA,
for their excellent support during launch, flight, and payload recovery.
This study was supported by
a Grant-in-Aid for Scientific Research
and a Grant-in-Aid from the Global COE program
``Next Generation Physics, Spun from Universality and Emergence''
from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) of Japan,
and the Japan Society for Promotion of Science for Young Scientists.
|
\section*{Introduction and results}
The implications of the violation of Bell inequalities \cite{Bell64}
in Quantum Mechanics (QM) are still controversial.
On one side, their logical and probabilistic nature has been
recognized and discussed \cite{Fine}, \cite{GM},
\cite{Pit89} and
their fundamental meaning traced back \cite{Pit94}
to Boole's general notion
of \textit{conditions of possible experience} \cite{Boole}.
In this perspective, their violation, which excludes representability
of quantum mechanical predictions by a classical probability
theory, shows general and fundamental differences between
the classical and the quantum notion of event.
On the other, the violation of Bell inequalities is interpreted
by many authors in terms of \lq\lq non local properties\rq\rq\ of QM;
an extensive review of such positions is contained in Ref. \cite{Griff}
and a discussion of their basic logical steps is in
Refs. \cite{Hardy}, \cite{Stapp1}, \cite{Mermin}, \cite{Stapp2}.
In our opinion, in general, the use of Bell inequalities for the
analysis of QM and of its interpretation is in a sense incomplete
because their violations simply amounts to a negative result,
i.e. the inconsistency between QM and \textit{the whole set}
of assumptions entering in their derivation.
The present paper is an attempt to reconsider the situation from a more
constructive point of view, beginning with QM predictions, trying to
represent them by classical probability models in a sequence of steps
and asking which step may fail.
We recall that adopting a classical probability model
exactly amounts to assuming that \textit{all variables together
take definite, even if possibly unmeasurable, values}, with
definite probabilities.
Since a classical probability model is equivalent to its set of
predictions for \textit{all} correlations, while QM only predicts
correlations between \textit{compatible} observables,
classical representability also amounts to an extension of QM
predictions to \textit{unmeasurable correlations}, e.g.,
correlations between two components of the spin
\textit{of the same particle},
or polarization of \textit{the same photon} along different directions.
A central role is therefore played by the notion of
\textit{extension of QM predictions to unmeasurable correlations}.
We shall use basic and general constraints
on such \textit{extensions} to discuss the
violation of Bell inequalities in terms of actual properties of
existing (partial) extensions rather than in terms of incompatibility
between assumed principles.
In particular, this will allow for a definite answer to the question
\textit{what precisely is \lq\lq influenced\rq\rq } in arguments on
non-local effects in QM.
The extension problem outlined above has been studied in general in \cite{B-M}.
One of the result (see Sect.1) is that classical
representability \textit{always holds} for yes/no observables
with compatibility relations described by \textit{a tree graph}, i.e. a graph
in which any two points are connected by exactly one path,
with points representing observables and links predicted correlations.
In the case of four yes/no observables, $A_1,A_2,B_1, B_2$,
$A_i$ compatible with $B_j$, this implies that, for fixed $B$,
any pair of probabilities for the subsystems
$A_i, B$ $i = 1,2$, assigning the same probability to the
outcomes of $B$, admits extensions to a probability
on the whole system $A_1,A_2,B$, each giving a definite value to the
possibly unmeasurable correlation $\langle A_1 A_2\rangle$.
As we will see, the results on tree graphs also imply that
\textit{if a value can be assigned to} $\langle A_1 A_2\rangle$
consistent with two choices $B_1$, $B_2$ for $B$,
then the whole system $A_i, B_j$ admits a classical representation.
Since the converse is obvious, it follows
that the existence of a classical representation for given set of
expectations $\langle A_i\rangle $, $\langle B_i\rangle$ and
correlations $\langle A_i B_j\rangle$
is equivalent to the possibility of giving a value to
$\langle A_1 A_2\rangle$ which is
\textit{consistent with the two choices for $B$}.
The analysis of QM models with four yes/no observables is therefore reduced
in general to the consistency (i.e. a non void intersection)
between the ranges of the (hypotetical, unmeasurable) values
for the correlation $\langle A_1 A_2\rangle$ allowed by (measurable) given
correlations of $A_1$ and $A_2$ with $B_1$ and $B_2$.
With respect to the same conclusion obtained by Fine \cite{Fine}, we
stress that probabilistic descriptions of three-observable
subsystem \textit{automatically exist} and only their compatibility
is in question; moreover, only a repeated application of the results for
tree graphs is required by our argument.
In Sect.2 , the possible range for $\langle A_1 A_2\rangle$,
depending on the measurable correlations $\langle A_i B\rangle$,
is characterized in terms of elementary three-observable
inequalities. QM states which violate Bell inequalities
are shown to give rise to disjoint intervals for the
admissible values of $\langle A_1 A_2\rangle$
in automatically existing probabilistic models for $A_1,A_2,B_1$
and $A_1,A_2,B_2$.
This clearly shows that classical representability of
such states exactly fails in the attribution of a value to an
unmeasurable correlation.
Actually, since BCH inequalities, more precisely, the
eight inequalities discussed by Fine \cite{Fine},
are equivalent to classical representability,
their violation exactly amounts to the inconsistency of such
an attribution.
The same discussion and result apply to the states introduced, on the
same set of observables, by Hardy \cite{Hardy} and
exploited in Stapp's work \cite{Stapp1}, discussed by Mermin \cite{Mermin}.
In particular, Stapp's and Mermin's discussion exactly concerns
partial extensions and the origin
of the inconsistency of the ranges of value for
an unmeasurable correlation $\langle A_1 A_2\rangle$
associated as above to the correlations of $A_i$
with $B_1$ and $B_2$ respectively.
In Sect.3, the experiment proposed by Greenberger, Horne,
Shimony and Zeilinger (GHSZ) \cite{GHSZ} is analyzed along the same lines.
The QM correlations defined by the GHSZ state,
between observables $A_i$, $B_i$, $C_i$, $i = 1,2$,
are shown to extend to \textit{all correlations} within the sets of observables
$A_i, B_i, C_1$ and $A_i, B_i, C_2$; both extensions give
\textit{unique values} for an \textit{ unmeasurable correlation},
$\langle A_1 A_2 B_1 B_2 \rangle = \pm 1$, and the sign depends
on the choice between $C_1$ and $C_2$.
Again, only the attribution of a value to an unmeasurable correlation
depends on the choice of an additional observable.
In the last Section we will comment on the fact that such a dependence
is the only basis for arguments about non local and non causal effects.
\section{\label{sec:ext}Extensions of QM predictions}
A basic notion is that of yes/no observables, denoting physical devices
producing, in each measurement, i.e. in each application to a physical system,
two possible outcomes; by \textit{experimental setting} we will denote a
collection of yes/no observables.
Within an experimental setting, an \textit{(experimental) context} will
denote a (finite) set of observables
which \lq\lq can be measured together
on the same physical system\rq\rq, in the precise sense that
their joint application to a physical system
gives rise to a statistics of their joint
outcomes given by a classical probability,
i.e., by a normalized measure on the
Boolean algebra freely generated by them.
Observables appearing together in some context will be called
\textit{compatible}; their correlations will be called
\textit{measurable}, or \textit{observable}.
In the following, Quantum Mechanics will be interpreted as a theory
predicting probabilities associated to experimental contexts defined
by sets of commuting projections.
Joint statistics of non compatible observables are not defined in the above
setting; correlations between incompatible observables will be called
\textit{unmeasurable}, or \textit{unobservable}.
In particular, we will consider the
\textit{Bell-Clauser-Horne (BCH) experimental setting},
consisting of two pairs ($A_1$, $A_2$), ($B_1$, $B_2$), of incompatible
\textit{yes/no} observables, taking values in $\{ 0,1 \}$;
the $A_i$ are compatible with the $B_j$,
i.e. they can be measured together, in the sense introduced above.
Such variables may be interpreted in terms of
spin components or photon polarizations, each pair referring to one
of a pair of particles, possibly in space-like separated regions.
The experimental contexts given by QM consist, in the BCH setting,
of the four pairs $\{A_i, B_j\}$; probabilities are associated to
contexts, see Proposition \ref{prop:mis} below, by the observed relative
frequencies $\langle A_i\rangle$, $\langle B_j\rangle$,
and $\langle A_i B_j\rangle$.
QM predictions for the BCH setting precisely consist
in probabilities on contexts,
given by the spectral representation for the
corresponding sets of commuting projectors.
Such a \textit{partial} nature of QM predictions is a very general fact since
the (ordinary) interpretation of QM precisely consists of a set of probability
measures, each defined, by the spectral theorem,
on the spectrum of a commutative
subalgebra of operators in a Hilbert space.
Such a structure has been formalized in \cite{B-M} as a
\textit{partial probability theory} on a \textit{partial Boolean algebra}.
A basic problem for the interpretation of QM is \textit{the necessity} of such
partial structures; in other words, \textit{whether they admit a classical
representation}, i.e. common probabilistic classical description in
a probability space
$(X,\Sigma, \mu)$, where $X$ is a set, $\Sigma$ a $\sigma$-algebra of
measurable (with respect to $\mu$) subsets and $\mu$ is a probability
measure on $X$ and all (yes/no) variables are described
by characteristic functions of measurable subsets.
Clearly, such a notion covers all kinds of \lq\lq hidden variables\rq\rq\
theories, all ending in the attribution of values to all the observables, with
definite probabilities.
Extensions have been discussed in general in \cite{B-M}. Non trivial
extensions, describing QM predictions through a reduced number of contexts,
have been shown to arise in some generality and some of the results
can be described in terms of graphs:
\begin{prop}\label{prop:tree1}
Consider any set of probabilities $p_i$ on a set of yes/no
observables $A_i$ and correlations $p_{ij}$ on a subset of pairs $A_i, A_j$,
defining a probability on each pair
$A_i, A_j$, with $ p_i = \langle A_i \rangle$,
$ p_j = \langle A_j \rangle$, $ p_{ij} = \langle A_i A_j \rangle$;
describing observables $A_i$ as points and the above pairs as links
in a graph, any subset of predictions associated to \textit{a tree subgraph}
admits a classical representation.
\end{prop}
\begin{prop}\label{prop:tree2} the same holds with
\textit{yes/no observables} $A_i$ substituted by
free Boolean algebras $ \mathcal{A}_i $,
$p_i$ by probabilities on
$ \mathcal{A}_i $, $p_{ij} $ by probabilities on the Boolean algebra freely
generated by the union of the sets of generators of
$ \mathcal{A}_i $ and $ \mathcal{A}_j $.
\end{prop}
Propositions \ref{prop:tree1} and \ref{prop:tree2} are proven in \cite{B-M}
by induction on the number of links of the correlation tree, by an explicit
construction of a probability on the Boolean algebra generated by two
algebras in terms of conditional probabilities
with respect to their intersection. The resulting probabilities are
in general not unique.
The above Propositions can be interpreted as providing automatic
conditions for the realization of the possibility advocated by
Einstein, Podolsky and Rosen \cite{EPR} to attribute a value to
correlations between incompatible observables consistently
with the QM predictions for the remaining, measurable, correlations;
actually, they do not make reference to QM and give conditions
on compatibility relations allowing the extension of any set of
predictions.
It is also useful to recall the following elementary fact \cite{B-M},
which implies that it is sufficient to analyze correlations since they
completely define a probability measure.
\begin{prop}\label{prop:mis}
In a classical representation for $n$ observables $A_1,\ldots ,A_n$, the
probability measure is completely defined by all expectation values
$\langle A_i \rangle$ together with all possible correlations
$\langle A_i A_j\rangle$,
$\langle A_i A_j A_k\rangle$,\ldots, $\langle A_1\ldots A_n\rangle$.
\end{prop}
In the BCH setting, the two subsets
$\{A_1, A_2, B_1\}$, $\{A_1, A_2, B_2\}$, with the correlations
$\langle A_j B_i \rangle$ $j=1,2,$
are described by tree graphs (see Fig.1 (a),(b))
and admit
therefore classical representations.
It follows that all QM predictions can be reproduced by a pair of
classical probability theories, one for
the Boolean algebra freely generated by
$A_1, A_2, B_1$ the other for $A_1, A_2, B_2$;
both sets of variables include $A_1$ and $ A_2$ and give
rise to predictions for the unmeasurable
correlation $\langle A_1 A_2 \rangle$.
In other terms, \textit{given a choice for} $B$, one
may consistently speak of the correlation $\langle A_1 A_2 \rangle$.
Clearly, the existence of a classical representation for all the predictions
in the BCH setting implies the existence of a value for
$\langle A_1 A_2 \rangle$ compatible with
the different choices of $B$.
On the other hand, if it is possible to give a value to
$\langle A_1 A_2 \rangle$
independently of the choice $B_1$, $B_2$ for $B$, this defines,
by Proposition \ref{prop:mis}, a
probability on the Boolean algebra $\mathcal{A}_{12}$ generated by $A_1, A_2$;
then, the application of Proposition \ref{prop:tree2}
to $\mathcal{A}_{12}$, $\{B_1 \}$, $\{B_2 \}$
implies classical representability of all the predictions for the
BCH setting (see Fig.1 (c)).
\begin{figure}
\begin{center}
\includegraphics[width=0.7\textwidth]{graph.png}
\caption{(a) BCH compatibility graph;
(b) two tree subgraphs; (c) tree graph corresponding to
extension of the measure on the algebra generated by $A_1,A_2$ }
\end{center}
\end{figure}
Classical representability and BCH inequalities
(in the sense of Ref.\cite{Fine}) are therefore
equivalent to the possibility of attributing a common value to the
unmeasurable correlation $\langle A_1 A_2 \rangle$
in classical models for $A_1, A_2, B_1$ and $A_1, A_2, B_2$.
In the next Section, we shall compute explicitly
the range of $\langle A_1 A_2 \rangle$
allowed by the QM predictions for a pair of spin variables, in the zero total
angular momentum state and for the state introduce by Hardy \cite{Hardy}.
\section{\label{sec:constr}Explicit Constraints
for systems of three and four observables}
In this section we first recall simple constraints, for three
$ \{ 0, 1 \} $ valued random variables $ A_1, A_2, B $,
satisfied by $\langle A_1 A_2\rangle$ given
$\langle A_i \rangle$, $\langle B \rangle$ and
$\langle A_i B \rangle$. We then show that, for the
predictions given by both the QM states discussed
by Bell and Hardy \cite{Hardy}, such constraints give
rise, for different choices of $B$, to incompatible
values for $\langle A_1 A_2\rangle$.
The possible ranges of correlations in a classical model
for $\{A_1,A_2,B\}$ are given by the Bell-Wigner
polytope \cite{Pit89}, also discussed in \cite{B-M}.
We only need a subset of the corresponding set of
inequalities, namely
\begin{equation}
\label{P1}
0 \leq \langle A_1 A_2\rangle \leq \langle A_1 \rangle
\end{equation}
\begin{equation}
\label{P2}
\langle A_1 B\rangle + \langle A_2B\rangle -\langle B \rangle
\leq \langle A_1 A_2 \rangle
\leq \langle A_2 \rangle - \langle A_2B \rangle +\langle A_1B \rangle
\end{equation}
\begin{equation}
\label{P3}
\langle A_1\rangle +\langle A_2\rangle +\langle B\rangle -
\langle A_1B \rangle -\langle A_2 B\rangle - 1
\leq \langle A_1 A_2\rangle
\end{equation}
Eq. (\ref{P1}) is trivial; the first relation in eq. (\ref{P2})
immediately follows from
$ \langle (A_1 - B)(A_2 - B) \rangle \geq 0$, which holds
since for $B=1$ both factors are non positive, for $B=0$
they are both non negative; moreover, the first relation
in eq. (\ref{P2}) implies the second by interchanging $B$ and $A_2$
and eq. (\ref{P3}) by interchanging $B$ and $1-B$.
The complete set of constraints \cite{Pit89} includes three
additional inequalities; they are obtained from the above relations
through the interchange of $A_1$ and $A_2$
and the substitution of $B$ with $1$ in eqs. (\ref{P1}), (\ref{P2})
(and are not relevant for our analysis).
The constraints given by eqs. (\ref{P1}), (\ref{P2}) (\ref{P3}),
applied to QM predictions for mean values and \textit{observable}
correlations, may give rise to ranges of values for the
unobservable correlation $ \langle A_1 A_2 \rangle $ which are disjoint
for different choices of $B$.
Consider in fact a system of two spin $1/2$ particles in the singlet state
${\frac{1}{\sqrt{2}}(\lvert+-\rangle- \lvert-+\rangle)}$
and associate $A_1$ and $A_2$ with the
projectors on the spin up state along directions $\hat{x}$ and
$\hat{z}$ for the first particle and $B_1$ with the projector on
the spin down state along the direction
$\frac{\hat{x}+\hat{z}}{\sqrt{2}}$ for the second particle, namely
\begin{equation}
\frac{1+\sigma_x}{2},\qquad \frac{1+\sigma_z}{2},
\qquad \frac{1}{2}-\frac{\tau_x+\tau_z}{2\sqrt{2}},
\end{equation}
where $\sigma_i$'s and $\tau_i$'s are Pauli matrices acting,
respectively, on the first and on the second subsystem.
Quantum mechanical predictions give
\begin{align}\label{eq:val1}
\langle A_1\rangle = \langle A_2\rangle = \langle B_1\rangle =
\frac{1}{2} \ ,\\
\label{eq:val13} \langle A_1 B_1 \rangle =
\langle A_2 B_1\rangle = \frac{1}{4} +\frac{\sqrt{2}}{8} \ ,
\end{align}
Denoting by $\langle\text{ }\rangle_1$ correlations in
a probabilistic model for $\{A_1,A_2,B_1\}$ eqs. (\ref{P1}) and (\ref{P2})
give
\begin{equation}\label{eq:rng1}
\frac{1}{4} < \frac{\sqrt{2}}{4}\leq
\langle A_1 A_2\rangle_1\leq \frac{1}{2} \ .
\end{equation}
If $B_1$ is substituted by $B_2$, representing the projector
\begin{equation}
\frac{1}{2}-\frac{\tau_x-\tau_z}{2\sqrt{2}} \ ,
\end{equation}
QM predictions give $\langle B_2 \rangle=\frac{1}{2}$ and
\begin{equation}
\label{eq:val13'} \langle A_1 B_2\rangle =\frac{1}{4}+\frac{\sqrt{2}}{8} \ ,
\quad \langle A_2 B_2\rangle = \frac{1}{4}-\frac{\sqrt{2}}{8} \ .
\end{equation}
Eqs. (\ref{P1}) and (\ref{P2}) then give, for the unmeasurable
correlation $\langle\text{ }\rangle_2$ in
probabilistic models for $\{A_1,A_2,B_2\}$
\begin{equation}\label{eq:rng2}
0\leq \langle
A_1 A_2 \rangle_2\leq\frac{1}{2}-\frac{\sqrt{2}}{4} < \frac{1}{4} \ .
\end{equation}
It follows that the value of $\langle
A_1 A_2\rangle$ must belong to one of two
\textit{disjoint intervals}, depending
on whether one is considering the probability space for $\{A_1,A_2,B_1\}$
or that for $\{A_1,A_2,B_2\}$, i.e. whether one chooses to measure $B_1$ or
$B_2$ on the second subsystem.
Therefore, the validity of QM predictions for compatible observables
\textit{does not allow} to assign a definite value to the
correlation $\langle A_1 A_2\rangle$, independent of the choice of the
measurement to be performed on the second subsystem.
The second example is Hardy's experiment \cite{Hardy},
also exploited in Stapp's \cite{Stapp1} and Mermin's \cite{Mermin}
discussion; we shall refer to Mermin's notation.
As in the previous example, we have two spin $1/2$ particles and four yes/no
observables, denoted by $L1,L2$ and $R1,R2$, acting respectively on a
``left'' and a ``right'' particle.
The state is described, up to a normalization
factor, by the vector
\begin{equation}
\lvert \Psi\rangle = \lvert L1+,R1-\rangle -
\lvert L2-, R2+\rangle \langle L2-, R2+\lvert L1+, R1-\rangle\ ,
\end{equation}
where, e.g, $\lvert L1+,R1-\rangle$
indicates a simultaneous eigenstate of the commuting
observables $L1$ and $R1$ with eigenvalue $1$ on the left and $-1$ on the right.
The variables $A_i$, $i=1,2$, are associated with the propositions
``the result of the measurement of $Ri$ is $+1$'', the same
for $B_i$ and $Li$; as before, $A_i$ and $B_i$ take values in $\{0,1\}$.
Independently of the specific expressions for $Li$ and $Ri$,
QM predicts the following correlations
(corresponding to eqs.(6)-(9) in Ref. \cite{Mermin}):
\begin{eqnarray}
\label{eq:s1} 1-\langle A_2 \rangle -\langle B_1
\rangle+\langle A_2 B_1 \rangle = 0 \ ,\\
\label{eq:s2} \langle A_2 \rangle -\langle A_2 B_2 \rangle = 0 \ ,\\
\label{eq:s3} \langle A_1 B_2\rangle = 0 \ ,\\
\label{eq:s4} \langle A_1 \rangle - \langle A_1 B_1 \rangle > 0 \ ,
\end{eqnarray}
In eq. (\ref{eq:s4}), $>0$
is equivalent to $\neq 0$ in Ref. \cite{Mermin} since
$\langle A_1\rangle\geq \langle A_1 B_1 \rangle$ always holds.
Eqs. (\ref{P3}), (\ref{eq:s1}) and (\ref{eq:s4}) give,
in all classical models for $\{A_1,A_2,B_1\}$
\begin{equation}
\langle A_1 A_2\rangle_1 \geq \langle A_1\rangle +
\langle A_2\rangle +\langle B_1\rangle -\langle A_1B_1
\rangle -\langle A_2 B_1\rangle -1 = \langle A_1\rangle -\langle A_1B_1
\rangle > 0 \ ;
\end{equation}
on the other hand, in classical models for $\{A_1,A_2,B_2\}$,
eqs. (\ref{P2}), (\ref{eq:s1}) and (\ref{eq:s4}) give,
\begin{equation}
\langle A_1 A_2\rangle_2 \leq \langle A_2\rangle -\langle A_2B_2\rangle +
\langle A_1B_2 \rangle = 0 \ ,
\end{equation}
which also implies, by eq. (\ref{P1}),
\begin{equation}
\langle A_1 A_2\rangle_2 = 0 \ .
\end{equation}
As in the previous example, no value for the unmeasurable
correlation $\langle A_1 A_2\rangle$ is compatible
with different choices of $B_i$.
A peculiarity of this case is that in classical models for
$\{A_1,A_2,B_2\}$ the correlation $\langle A_1 A_2\rangle$ is completely fixed
(to $0$) by the constraints imposed by measurable correlations.
Attributing the $0$ value is actually equivalent to asserting the logical
implication $A_1 \rightarrow A_2^c$, i.e.
``\textit{if} the measurement of $R1$ gives $+1$,
\textit{then} the measurement of $R2$ gives $-1$''; such an implication
\textit{holds in all classical models for}
$\{A_1,A_2,B_2\}$ \textit{and fails
in all classical models for} $\{A_1,A_2,B_1\}$.
\section{\label{sec:GHSZ}The perfect correlation model of
GHSZ}
In this Section, the above analysis is applied to the case considered
in \cite{GHSZ} and discussed in \cite{Mermin2} within a general
framework unifying Bell and Kochen-Specker kinds of results.
With a slight modification of the above terminology,
the experimental setting involved in the GHSZ model consists
of six yes/no observables, which can be described as $A_i$, $B_i$, $C_i$,
$i = 1,2$,
taking values $1$ and $-1$; equivalently, by the propositions
$P_i$, $Q_i$, $R_i$ respectively asserting
$A_i = 1$, $B_i = 1$, $C_i = 1$.
The experimental contexts, i.e. the sets of compatible observables
are given by the possible choices of (at most) one $A$, one $B$ and one $C$.
In Refs. \cite{GHSZ} and \cite{Mermin2} observables associated to
different letters are interpreted as measured in distant,
space-like separated regions.
The above observables reproduce the example of Ref. \cite{GHSZ}, Sect. III,
with the identification $ A_i = A(\phi_i)B(0)$, $B_i = C(\phi_i)$,
$C_i =D(\phi_i)$, with $\phi_1 = 0$ $\phi_2= \pi/2$;
in Mermin's spin notation, they should be read as
$A_1 = - \sigma^1_x$, $A_2 = - \sigma^1_y$,
$B_1 = \sigma^2_y$, $B_2 = \sigma^2_x$,
$C_1 = \sigma^3_y$, $C_2 = \sigma^3_x$.
In general, quantum mechanical predictions consist in probability
assignments within each of the eight contexts defined by the choice of
three indexes $i$; e.g. a context is given by the choice
$A_1$, $B_1$, $C_1$, and the associated predictions by a probability on the
Boolean algebra freely generated by the propositions
$P_1$, $Q_1$, $R_1$.
The crucial point of GHSZ and Mermin's analysis is the observation that, for
suitable states, QM predictions give \lq\lq perfect correlations\rq\rq\
(each correlation involving observables in a fixed context),
which are not compatible with any
(context-independent) assignment of values to the variables
$A_i$, $B_i$, $C_i$.
In fact, the state considered by GHSZ and Mermin can be written, for the
Mermin spin variables $\sigma^i_k$,
in the usual notation referring to
the eigenvalues of $\sigma^i_3$,
\begin{equation}
\label{psi}
\lvert \Psi \rangle = \frac{1}{\sqrt 2} \ (\lvert +++\rangle - \lvert---\rangle)
\end{equation}
and gives the correlations
\begin{eqnarray}
\label{perfcorr}
\langle A_1 B_1 C_1 \rangle = -1\ , \qquad \langle A_2 B_2 C_1\rangle
= -1\ , \\
\langle A_2 B_1 C_2 \rangle = -1\ , \qquad
\langle A_1 B_2C_2\rangle = +1\ . \label{perfcorr2}
\end{eqnarray}
Values $a_i, b_i, c_i = \pm 1$ assigned to the above variables and
reproducing the above correlations would satisfy the relations
\begin{equation}
a_1 b_1 c_1 = -1\ , \ \ a_2 b_2 c_1= -1\ , \ \ a_2 b_1 c_2 = -1\ ,
\ \ a_1 b_2 c_2 = +1\ ,
\end{equation}
which have no solution since their product gives $1$ for the l.h.s.,
$-1$ for the r.h.s..
Following the program outlined above and using the notions
and results of Sect. 1,
we will extend the GHSZ and Mermin argument to show:
\begin{itemize}
\item[$i)$] the correlations eqs. (\ref{perfcorr}), (\ref{perfcorr2})
\textit{and all the other QM
predictions} given by the GHSZ state, eq. (\ref{psi}),
extend to classical probabilities $p_1$, $p_2$
on the \textit{two} subsystems defined by
$\{A_1, A_2, B_1, B_2, C_1\}$ and $\{A_1, A_2, B_1, B_2, C_2\}$
\item[$ii)$] unique and different values, $\pm1$, are given by such
$p_1$ and $p_2$ to the correlation $\langle A_1 A_2 B_1 B_2 \rangle $,
all the other correlations between the $A$ and $B$ observables
admitting common values;
\item[$iii)$] pairs of probabilities on the two subsystem
introduced in $i)$ extend to a probability on
the entire system generated by
$A_i, B_j, C_k$, $i,j,k, = 1,2$,
exactly if they coincide on the subsystem generated by
$A_1,A_2,B_1,B_2$.
\end{itemize}
Again, the conclusion is that \textit{exactly a non observable
correlation}, $\langle A_1 A_2 B_1 B_2 \rangle $, \textit{depends} on the
choice of the observable $C_i$.
In order to derive $i)$ - $iii)$,
we first observe that $iii)$ follows immediately
from Proposition \ref{prop:tree2}, applied to the Boolean algebra generated by
$P_i, Q_j$ and to its correlations with the observables
$R_1$ and $R_2$, forming a tree graph.
Concerning $ii)$, observe that eqs. (\ref{perfcorr}) imply,
for any probabilistic model reproducing them,
$A_1 B_1 C_1 = -1 $ and $ A_2 B_2 C_1 = -1 $ with probability $1$,
so that, with probability $1$,
$ A_1 B_1 A_2 B_2 = 1 $; equivalently,
$ \langle A_1 B_1 A_2 B_2 \rangle = 1 $.
In the same way, eqs. (\ref{perfcorr2}) imply
$ \langle A_1 B_1 A_2 B_2 \rangle = - 1 $; such relations hold therefore
for all probabilistic models reproducing, respectively, the QM correlations
of the $A,B$ observables with $C_1$ and $C_2$. The existence of probabilities
giving common values to all the other correlations follows from the
construction below.
The most involved issue is $i)$, which is non trivial since quantum mechanical
predictions involve eight different contexts and $i)$ states that the
predictions given by the state (\ref{psi}) extend
to \textit{all correlations} between variables inside each one of
\textit{only two} contexts.
Notice that Propositions \ref{prop:tree1} and \ref{prop:tree2}
do not imply such an extension, since
the corresponding set of correlations is not given by a tree graph and in fact
not all quantum states admit it, as also implied by the
BCH analysis; the exact form of the quantum predictions we are going
to extend is therefore important. It is easy to see that the state given
by eq. (\ref{psi}) gives rise to the correlations
\begin{equation}
\label{otherpred1}
\langle A_i \rangle = \langle B_i \rangle = \langle C_i \rangle =
\langle A_i B_j \rangle = \langle A_i C_j\rangle = \langle B_i C _j
\rangle =0 \ , \ i,j=1,2
\end{equation}
\begin{equation}
\label{otherpred2}
\langle A_1 B_2 C_1 \rangle = \langle A_2
B_1 C_1 \rangle = \langle A_2 B_2 C_2 \rangle= \langle A_1 B_1 C_2
\rangle = 0 \ .
\end{equation}
We have to show that the all the correlations given by eqs.
(\ref{perfcorr}),(\ref{perfcorr2}),(\ref{otherpred1}),(\ref{otherpred2})
involving
the observables $A_i, B_j, C_1$
can be represented by a classical model,
and the same for those involving $A_i, B_j, C_2$, and that the two models
give the same correlations between variables
$A_i, B_j$ with the only exception of
$\langle A_1 A_2 B_1 B_2 \rangle $.
The first classical model is defined as follows: let $A_i, C_1$
denote variables
taking values $1$ and $-1$ with probability $1/2$, i.e.
\begin{equation}
\label{mod11}
A_i^2 = 1\, ,\ C_1^2 = 1 \ , \ \
\langle A_i \rangle = \langle C_1 \rangle = 0 \ , \ \
\langle A_1 A_2 \rangle = \langle A_i C_1 \rangle = 0
\end{equation}
and define
\begin{equation}
\label{mod12}
B_i \equiv - A_i C_1 \ .
\end{equation}
Eqs. (\ref{mod11}) and (\ref{mod12}) immediately imply eqs. (\ref{perfcorr})
and all eqs. (\ref{otherpred1}),(\ref{otherpred2})
not involving $C_2$.
The second model is defined in the same way, by independent variables
$A_i$ and $C_2$ satisfying the same relations as in eqs. (\ref{mod11})
and by variables $B_i$ now defined as
\begin{equation}
\label{mod22}
B_1 \equiv - A_2 C_2 \ , \ \ B_2 \equiv A_1 C_2 \ .
\end{equation}
Eqs. (\ref{perfcorr2}) and all eqs. (\ref{otherpred1}), (\ref{otherpred2})
not involving $C_1$ immediately follow.
All the correlations defined by the two models between $A$ and $B$ variables
follow from eqs. (\ref{mod11}), with $C_2$ replacing $C_1$ for the second model,
(\ref{mod12}),(\ref{mod22}); in both models
$\langle A_1 A_2 \rangle = 0 $ by definition,
\begin{equation}
\label{modsother}
\langle B_1 B_2 \rangle = \pm \langle A_1 A_2 \rangle = 0
\end{equation}
and
$$
\langle A_i B_1 B_2 \rangle = 0 = \langle B_i A_1 A_2 \rangle \ ,
$$
both expectations reducing to $ \pm \langle A_k \rangle $ for some $k$.
The two models give therefore identical predictions for all the correlations
between the $A$ and $B$ variables, with the only exception
\begin{equation}
\label{modsdiff}
\langle A_1 A_2 B_1 B_2 \rangle = \pm 1 \ .
\end{equation}
\section{\label{conclusions}Conclusions}
The discussion of Bell inequalities usually begins with their
derivation from general \textit{locality} and \textit{reality}
principles. Reality principles \textit{include} hypothetical
common attributions of values to observables even when they are not
compatible according to QM; once such attributions are assumed,
any logical or probabilistic consideration automatically concerns
extensions of QM predictions to unmeasurable correlations between
incompatible observables.
Moreover, arguments about the possibility of considering
measurements which \textit{could have been performed}
and the assumption that the results of such \lq\lq unperformed\rq\rq\
measurements satisfy the predictions of QM for their correlation with
other, compatible and performed, measurements, lead
\cite{Stapp1} \cite{Mermin}
to logical or probabilistic relations
between quantum mechanically incompatible observables
which precisely amount
to partial extensions of QM predictions.
We have shown that, independently of any additional principle
or argument, partial extensions of QM predictions are
always possible in cases described by tree graphs,
that the ranges of values for correlations between
quantum mechanically incompatible observables obtained by
such extensions depend in general on the considered set of observables
and that they may be incompatible for different sets.
As shown by the Hardy and the GHSZ examples, this may happen
even when \textit{all extensions},
to a fixed allowed set of observables,
of a set of QM predictions give the same
\textit{perfect correlations} to quantum mechanically incompatible
observables.
If the \textit{dependence of such correlations on a set
of observables} allowing for a partial extension
is not taken into account, the result appears as a
contradiction between logical consequences of QM predictions
(and of experimental results).
Moreover, precisely unmeasurable correlations depend on
a choice in a far or future space-time regions in the discussion
of locality and causality principles.
In fact, in the BCH setting, identifying $B_i$ with observables measured
in a space-like separated region, only and exactly
$\langle A_1 A_2 \rangle $ depends on the choice of $B_i$;
alternatively, if $B_i$ is measured in a future region,
such a dependence violates causality
in the precise sense that any hypotetical record of
the value of $\langle A_1 A_2\rangle$ depends on a future choice.
All the above \lq\lq violations\rq\rq\ exactly concern
extensions of QM predictions to \textit{correlations
between quantum mechanically incompatible observables}, e.g., between
two different spin components of \textit{the same} particle or polarization
directions of \textit{the same} photon.
A discussion of the violation of BCH inequalities in terms of
unobservable correlations has also been given in Ref. \cite{Cohen}.
A (different, but related) notion of \lq\lq extension of
quantum correlations\rq\rq\ is also central
in the recent discussion of the Einstein-Podolsky-Rosen notion of
reality in Ref. \cite{Nistico}.
|
\section*{Introduction}
Logarithmic Conformal Field Theories (LCFTs) are characterized by the fact that the dilatation operator
$L_0$ is non-diagonalizable and has a Jordan cell structure. This property leads to the appearance of logarithms
in correlation functions and to indecomposable operator product expansions. Such theories have emerged recently
in many physical situations such as geometrical problems, disordered free fermions models in 2+1 dimensions
or the AdS/CFT correspondence in string theory.
Geometrical applications include self-avoiding random walks (polymers) or percolation (see {\it e.g.} \cite{SaleurPoly,MathieuRidout}).
Although it would be fair to say to these problems are well understood, a fully consistent field theory
describing geometrical observables ({\it e.g.}, four point correlation functions in the bulk) is still missing. Another very exciting question,
with considerable applications
to condensed matter, concerns the critical point in non-interacting disordered fermion models in 2+1 dimensions, which
are believed to be described by $c=0$ LCFTs. Physical realizations of such systems are given by transition between
plateaux in the Integer/Spin Quantum Hall Effects. The appearance of logarithmic correlators in
such theories cannot be avoided \cite{Cardylog} and a deep understanding of LCFTs is necessary in order to describe
their low-energy physics. While much has been done about the Spin Quantum Hall Effect, as it can be related to the
classical percolation problem \cite{SQHE}, very little is known about the field theory describing the transition in
the Integer Quantum Hall Effect (see \cite{Zirnbauer} for a review). Logarithms in CFTs also arise naturally in the context
of non-linear sigma models with super target space. These quantum field theories play a major role in the AdS/CFT duality.
For example, the PSL($2|2$) sigma model is related \cite{strings} to strings living in $\mathrm{AdS}_3 \times S^3$.
The key feature of logarithmic CFTs, as opposed to simpler non-unitary theories (such as the Yang-Lee singularity, or other non-unitary
minimal models), is indecomposability.
This property was probably observed first by Rozansky and Saleur \cite{RozanskySaleur} who studied a WZW model with U($1|1$) supergroup
symmetry and vanishing central charge. They related the non-simplicity of U($1|1$) to the possibility
of non-diagonalizability of $L_0$ and logarithmic dependence in four-point correlation functions.
The study of logarithmic $c=0$ theories in a more systematic fashion then begun with a serie of papers
by Gurarie and Ludwig. Gurarie \cite{Gurarie1} first noticed that logarithmic operators were necessary in order to construct
a consistent field theory at $c=0$. A similar observation was made by Cardy using a replica approach \cite{Cardylog}.
Gurarie \cite{Gurarie2} and Gurarie and Ludwig \cite{GurarieLudwig} then related the existence of a logarithmic
partner $t(z)$ for the stress energy tensor $T(z)$ to ill-defined terms in operator product expansions. They introduced one of the
first\footnote{Several other indecomposability parameters were computed by Kausch and Gaberdiel a few
years earlier \cite{KauschGaberdiel}, in a different context.} indecomposability parameters, usually denoted $b$, which at the time
was interpreted as a new `anomaly' that
would play the role of a central charge when $c=0$. Using some heuristic arguments, they were also able to predict two possible
values, $b=-\frac{5}{8}$ and $b=\frac{5}{6}$, which would distinguish between two fundamentally different LCFTs.
Instead of a single parameter $b$, it is now well accepted that a LCFT is characterized rather by a complex
structure of indecomposable Virasoro modules, with a infinite number of indecomposability parameters needed to describe the whole pattern.
Two lines of thought have been considered. The first one is to deal directly with abstract indecomposable Virasoro modules
to try to understand and classify their structures \cite{Rohsiepe,KauschGaberdiel, MathieuRidout,MathieuRidout1,KytolaRidout}.
Progress has been steady in this direction, and the module involving $T$ and its partner $t$ at $c=0$ now appears
as a particular case. These algebraic studies have led to new predictions for indecomposability parameters for various
logarithmic pairs. The second idea is somehow more concrete, and consists in studying directly lattice models which
can be thought of as lattice regularizations of LCFTs. For specific values of the parameters, the lattice Hamiltonians
are non-diagonalizable and have a Jordan cell structure that mimics that of the continuum theory. This was mainly
done by Read and Saleur \cite{RS3}, and independently by Pearce, Rasmussen and Zuber \cite{PRZ,RP1,RP2}. This
approach essentially relies on the `similarity' between the indecomposable modules of the Temperley-Lieb algebra
and those of the Virasoro algebra. Both approaches yield a consistent picture of boundary (chiral) LCFTs, with
common algebraic structure and fusion rules deduced from both methods. As for the similarity, it is now better understood in
terms of common quantum group structures \cite{Azat}.
While the global structure of the Virasoro modules in the case of boundary LCFTs is slowly getting under control,
many questions remain about the existence and values of the parameters arising from indecomposability. It is,
moreover, tempting to think of these parameters as extensions
of the structure constants of the operator product algebra of the theory,
and thus to wonder what they encode physically, and whether it is possible to access them,
experimentally or at least numerically, in the context of lattice simulations.
Recently, Dubail, Jacobsen and Saleur \cite{DJS} suggested a concrete
method to measure {\sl one} indecomposability parameter: the $b$-number of Gurarie in the case of $c=0$ theories.
Their method involved a $c=0$ specific approach, the so called `trousers trick', and led to the measurement of $b$ for percolation and polymers.
The observed values---$b=-\frac{5}{8}$ and $b=\frac{5}{6}$ respectively, were found in agreement
with the predictions of Mathieu and Ridout \cite{MathieuRidout}.
The method of \cite{DJS} is geometrically very appealing, but unfortunately does not extend to many other LCFTs, nor
does it allow the study of indecomposable modules occurring at larger values of the conformal weight. It also does
not seem to be generalizable to the bulk (non-chiral) case. On the other hand, having at one's disposal a `probe'
to investigate the detailed structure of modules in LCFTs seems rather essential to make progress in this, so far,
very abstract topic. We have thus reconsidered the problem using a different route, which involves the identification
of a lattice stress energy tensor. The original idea for doing so goes back at least to a paper by Koo and Saleur \cite{KooSaleur},
who themselves generalized the pioneering work of Kadanoff and Ceva \cite{KadanoffCeva}. The upshot of the proposal in \cite{KooSaleur}
was that the Temperley-Lieb algebra does not only reproduce {\sl in finite size} indecomposable modules that mimic exactly several
indecomposables of the Virasoro algebra: on top of this, there exist (infinitely many) linear combinations of words in the Temperley-Lieb
algebra whose action in these modules reproduces, once properly interpreted, the action of all the Virasoro generators
{\sl in the thermodynamic limit}. It thus should be possible, in principle, to reconstruct from the lattice all states of the boundary
LCFT, and to measure all the matrix elements of all the Virasoro generators, hence to determine `experimentally' all the information
about the Virasoro modules in the model.
Of course, this program is very difficult to implement in practice. Thanks in part to the progress accomplished in \cite{RS3,DJS},
it is however not impossible, and this is what we do in this paper.
Once the general strategy is under control, it turns out that we can study many more cases than those considered so far in papers
on indecomposable Virasoro modules \cite{KauschGaberdiel, MathieuRidout1,KytolaRidout}.
These authors used mainly two different methods to compute indecomposability parameters. One is based on
the so-called Nahm-Gaberdiel-Kausch algorithm to compute fusion products between indecomposable Virasoro modules \cite{KauschGaberdiel};
while the other, used in Refs. \cite{MathieuRidout1,KytolaRidout}, consider different quotients of `glueing' of two Verma modules,
incomposability parameters are then obtained as a solution of singular-vector equations. Note also that they can be computed using
SLE considerations \cite{Kytola}, or even directly by solving the differential equations satisfied by several four-point
correlation functions. The latter method was one of the tools used by Gurarie and Ludwig \cite{GurarieLudwig2} to study the
case of the stress energy tensor at $c=0$. While it should be possible to generalize the methods of these standard references
to analyze the Jordan cells and calculate the indecomposability parameters in our cases, we find it more convenient
here to use a `short cut' analysis based on a generalization of the original arguments of Gurarie for the $c=0$ case. This is
discussed, as a preamble, in the first section. The general framework is then discussed in section 2, where we introduce the lattice models, and
use and generalize the algebraic arguments of Ref. \cite{RS3} to deduce general indecomposable structures. The next two sections are then
dedicated to numerical measurements of indecomposability parameters. All our results are consistent with the
predictions in \cite{KauschGaberdiel, MathieuRidout1,KytolaRidout}, but we obtain many results beyond these references---those
for the first few logarithmic minimal models $\mathcal{LM}(1,p)$ and $\mathcal{LM}(p,p+1)$ are summarized in table \ref{general_b_figure}.
A few conclusions are finally gathered in the last section.
\section{Preamble: indecomposability parameters and $\epsilon \rightarrow 0$ argument}
In this section, we define in simple terms the indecomposability parameter associated
with a general pair of logarithmic operators. We then extend the $c \rightarrow 0$
argument of Gurarie to other logarithmic CFTs and we show that this allows us to predict
the value of the indecomposability parameters of a given theory.
\subsection{Jordan cells and indecomposability parameters}
It is now well known that the appearance of logarithms in correlation function is related
to the non-diagonalizability of the $L_0$ operator \cite{Gurarie1}. This Jordan cell
structure of the Hamiltonian ($L_0$) is itself deeply related to the non-semisimplicity of
the underlying symmetry algebra of the theory. We will come back to these algebraic
considerations later and choose here to focus on the Jordan cell structure. Consider
a pair of logarithmic operators ($\phi(z)$,$\psi(z)$) with conformal weight $h$ that
are mixed by $L_0$ into a Jordan cell. In the basis ($\phi$,$\psi$), the generator of
the scale transformation reads
\begin{equation}
L_0 =
\left( \begin{array}{cc}
h & 1 \\
0 & h \end{array} \right).
\end{equation}
Invariance under global conformal transformations then fixes \cite{Gurarie1} the form
of the correlation functions
\begin{subequations}
\begin{eqnarray}
\left\langle \phi(z) \phi(0)\right\rangle &=& 0 \\
\left\langle \phi(z) \psi(0)\right\rangle &=& \frac{\beta}{z^{2h}} \\
\left\langle \psi(z) \psi(0)\right\rangle &=& \frac{\theta -2 \beta \log z} {z^{2h}} ,
\end{eqnarray}
\end{subequations}
where $\theta$ and $\beta$ are two parameters. While the constant $\theta$ is arbitrary
and can be canceled by a choice $\psi \rightarrow \psi - \frac{\theta}{2 \beta} \phi $, the parameter
$\beta$ is a fundamental number that characterizes the structure of the Jordan cell. It is
also important to remark that $\phi(z)$ must be a null-field by conformal invariance, that
is to say, introducing the usual Virasoro bilinear form, $\Braket{\phi|\phi} =0$.
Actually, we know more about the algebraic structure hidden behind the non-diagonalizability
of $L_0$. As we will see in details in the following, the fields $\psi(z)$ and $\phi(z)$ always
appear at the top and the bottom of a larger structure called a staggered Virasoro module \cite{MathieuRidout,Rohsiepe,KytolaRidout},
also called projective\footnote{Note that this nomenclature is somewhat dangerous as strictly speaking, these modules
may not be projective over the Virasoro algebra in the mathematical sense. However, they can be seen as
`scaling limit' of projective modules over the Temperley-Lieb algebra that arises in lattice models ({\it cf.} next section),
it is thus very tempting to call them projective anyway.} module in Ref \cite{RS3}. The fields
in these modules are organized in a diamond-shaped structure that we note
\begin{equation}
\mathcal{P}=
\begin{array}{ccccc}
&&\hskip-.7cm \psi &&\\
&\hskip-.2cm\swarrow&\searrow&\\
\xi &&&\hskip-.3cm \rho \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm \phi &&
\end{array}
.
\label{eq_general_proj}
\end{equation}
The arrows represent the action of Virasoro generators so that the whole structure can be
induced by action of the $L_n$'s on the field $\psi(z)$ while $\phi(z)$ belongs to an invariant
submodule. The whole module is reducible but indecomposable under the action of the Virasoro
algebra. Remark that $L_0$ maps $\psi$ onto $\phi$ due to the Jordan cell action. The conformal
weights of the different fields in~\eqref{eq_general_proj} satisfy
$h_{\xi} \leq h=h_{\phi}=h_{\psi} \leq h_{\rho}$. The field $\phi(z)$ is a null descendant of $\xi$
\begin{equation}
\displaystyle \phi(z) = A \xi(z), \ \ A = L_{-n} + \alpha_{(1)} L_{-n+1}L_{-1}+ \dots + \alpha_{(P(n)-1)} L_{-1}^n,
\label{eq_Aconv}
\end{equation}
where $n = h-h_{\xi}$ and $P(n)$ is the number of partitions of the integer $n$. The $\alpha_{(i)}$
coefficients are uniquely fixed by the null-vector condition $L_{+1} \phi = L_{+2} \phi =0 $.
If two fields $\psi(z)$ and $\phi(z)$ satisfy all these relations, we say that $\psi(z)$ is the
logarithmic partner of the null-field $\phi(z)$, with {\it indecomposability parameter} (also
called logarithmic coupling) $\beta$. Using the Virasoro scalar product, we see that
\begin{equation}
\displaystyle \beta = \Braket{\psi | \phi}.
\end{equation}
where we normalized $\xi(z)$ such that $\Braket{\xi | \xi} = 1$. Note that it is also possible
to define $\beta$ through the equation
\begin{equation}
\displaystyle A^\dag \psi (z) = \beta \xi (z).
\end{equation}
It is important to notice at this point that the choice that we adopted for the normalization of the operator $A$
is crucial for the value of $\beta$. Different choices have been used in the literature, and some of them may
yield simpler expressions for $\beta$. Unless otherwise indicated, we always use the convention given by eq.~\eqref{eq_Aconv},
which is consistent with the normalization of the stress energy tensor in $c=0$ theories: $T=L_{-2} I$ so $A=L_{-2}$
(see next paragraph).
Finally, let us remark that the field $\rho(z)$ never enters the computations of $\beta$ so that
it can be ignored as far as indecomposability parameters are concerned.
\subsection{$c \rightarrow 0$ catastrophe and the stress energy tensor}
\label{subsec_catastrophe}
We now show that
the $\beta$'s are fixed by a very simple argument relying on operator product
expansions (OPEs). The basic idea was developed in \cite{Gurarie1,GurarieLudwig}, and a similar discussion may be found in \cite{KoganNichols,Cardylog2}. Consider a general CFT with the following parametrization
of the central charge and of the Kac formula
\begin{subequations}
\begin{eqnarray}
\displaystyle c &=& 1 - \frac{6}{x (x+1)}, \\
\displaystyle h_{r,s} &=& \frac{ \left[(x+1)r - xs \right]^2 - 1}{4 x (x+1)}.
\end{eqnarray}
\end{subequations}
Conformal invariance fixes the OPE of an operator $\Phi_h(z)$ with itself to be of the form
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h}} \left[1 + \frac{2 h}{c} z^2 T(0) +\dots \right],
\end{equation}
where $T(z) = L_{-2} I$ is the stress energy tensor of the theory. This expression is
clearly ill-defined as $c \rightarrow 0$. In a nutshell, the idea of Gurarie was to
introduce another field $\Phi_{1,5}$ for percolation ({\it resp.} $\Phi_{3,1}$ for
dilute polymers) with conformal weight $h_t \equiv h_{1,5}=2$ ({\it resp.}
$h_t \equiv h_{3,1}=2$) at $c=0$ to cancel this divergence\footnote{Actually, the
original guess of Gurarie and Ludwig was $\Phi_{1,5}$ for dilute polymers and $\Phi_{3,1}$
for percolation. It shall become clear in the next sections that this result was correct
up to a switch, as already remarked in Refs. \cite{MathieuRidout, DJS} .}. Let us focus on the percolation case. When $c$ is slightly different
from $0$ ($x = 2 + \epsilon$), we can normalize\footnote{For the sake of clarity, we absorb the coefficient $C^{\Phi_{1,5}}_{\Phi,\Phi}/a_{\Phi}$ into
the normalization of $\Phi_{1,5}$ as this will play no role in the following. The coefficients $C^{k}_{i,j}$ are the usual structure constants that appear in OPEs; note also that $a_{\Phi} = C^{\Phi_{1,1}}_{\Phi,\Phi}$.} the field $\Phi_{1,5}$ such that the OPE reads
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h}} \left[1 + \frac{2 h}{c} z^2 T(0) + z^{h_t} \Phi_{1,5} (0)+ \dots \right].
\end{equation}
We then define a new field $t(z)$ as
\begin{equation}\label{eq_newfieldt}
\displaystyle \Phi_{1,5} (z) = \frac{2h \Braket{T | T}}{c \beta(\epsilon)} t(z) - \frac{2h}{c} T(z),
\end{equation}
where $\beta(\epsilon) = - \frac{ \Braket{T | T}}{h_{t}-2}$ and $\Braket{T | T}=\frac{c}{2}$. The OPE then involves quantities that are perfectly well defined as $c\rightarrow 0$
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h}} \left[1 + \frac{ h}{\beta} z^2( T(0) \log z + t(0)) + \dots \right],
\end{equation}
with $\beta=\lim_{\epsilon \rightarrow 0} \beta (\epsilon)$.
It is important to realize that the new field $t(z)$ is perfectly well defined as $c\rightarrow 0$, while the $L_0$ eigenvector $\Phi_{1,5}$
is not. In particular, one can then check
(see {\it e.g.} \cite{KoganNichols}) that the fields $T(z)$ and $t(z)$ then satisfy
the standard equations for logarithmic operators
\begin{subequations}
\begin{eqnarray}
\left\langle T(z) T(0)\right\rangle &=& 0 \\
\left\langle T(z) t(0)\right\rangle &=& \frac{\beta}{z^4} \\
\left\langle t(z) t(0)\right\rangle &=& \frac{\theta -2 \beta \log z}{z^4} ,
\end{eqnarray}
\end{subequations}
with $\theta$ a constant. A straightforward calculation using eq.~\eqref{eq_newfieldt} also shows that $L_{0} t = 2 t + T$ as
expected.
Note that it is quite general that one of the eigenvectors (here $\Phi_{1,5}$) of an operator (here $L_0$)
diverges as one tunes a parameter (here $\epsilon$) to approach an indecomposable point. One can then construct a new
Jordan vector (here $t$) by canceling the diverging part in the ill-defined eigenvector by taking an appropriate combination with the
eigenvector that has the same eigenvalue at the indecomposable point (here $T$).
$T(z)$ and its logarithmic partner $t(z)$ are a special
case of the general structure described in the previous paragraph. In particular,
they are organized in a diamond structure like~\eqref{eq_general_proj} with
$\xi = I$, $\phi = T$, $\psi = t$, $A = L_{-2}$ and $L_{2}t = \beta I$.
Gurarie and Ludwig then inferred the value of $\beta$ using
algebraic arguments along with some heuristic hypotheses \cite{GurarieLudwig}.
At this point, it is important to notice that it is also possible to compute $\beta$ using the simple limit process.
This was already noticed by Gurarie and Ludwig in the context of the replica approach \cite{GurarieLudwig2}.
Using the parametrization $x = 2 + \epsilon$, we find
\begin{subequations}
\begin{eqnarray}
\beta_{\mathrm{percolation}} &=& - \lim_{\epsilon \rightarrow 0} \frac{c/2}{h_{1,5}-2} = -\frac{5}{8}, \\
\beta_{\mathrm{polymers}} &=& - \lim_{\epsilon \rightarrow 0} \frac{c/2}{h_{3,1}-2} = \frac{5}{6},
\end{eqnarray}
\end{subequations}
which indeed are the expected values \cite{MathieuRidout, DJS}. In this sense,
it is possible to understand the structure of a logarithmic CFT as a limit of
non-logarithmic CFTs. We suspect that the values of indecomposability parameters
can be inferred in a similar fashion for any LCFT. To see this, we turn to a
slightly more complicated example.
\subsection{Generalization to other LCFTs and general formula for $\beta$}
Let us consider a generalization of this argument to a more complicated case.
We focus in this paragraph on the theory of symplectic fermions \cite{Kausch1,Kausch2}
that describes the scaling limit of dense polymers on the lattice. This theory has
$c=-2$ $(x=1)$, and is particularly simple as there is no interaction. The Jordan
cell structure can be understood in terms of free fermions, and the indecomposability
parameters can be, in principle, computed using this free fermion representation. This
theory was also considered in order to test the Nahm-Gaberdiel-Kausch algorithm
to compute fusion products between indecomposable Virasoro modules \cite{KauschGaberdiel}.
We choose here to focus on a concrete example: this theory is known to have a Jordan
cell at level 3 with equations
\begin{align}
L_0 \phi & = 3 \phi \\ \notag
L_0 \psi & = 3 \psi + \phi \\ \notag
\phi & = A \xi \\ \notag
A^\dag \psi & = \beta \xi \\ \notag
A &= L_{-1}^2 - 2 L_{-2},
\end{align}
where the parameter $\beta=\Braket{\phi|\psi}=-18$ is the logarithmic coupling
associated with this Jordan cell \cite{KauschGaberdiel}. Note that we chose a
different normalization convention for the operator $A$ than eq.~\eqref{eq_Aconv} in order
to match \cite{KauschGaberdiel}.
When we think of this theory as the continuum limit of a XX spin chain with
quantum $U_{q=i}(\mathfrak{sl}_2)$ symmetry with an even number of sites \cite{RS3}, the
conformal dimensions that appear in the spectrum are $h_{1,1+2j}$ with
$j \in \mathbb{N}$. We shall come back to the precise nature of the scaling
limit of such lattice models in the next section. All we need for what follows
is to identify the conformal weights of the fields $\xi$ and $\psi$ in the
spectrum. We find $h_{\xi} = h_{1,5} = 1$ and $h_{\psi} = h_{1,7} = 3$. It is
important to identify the operators in spectrum as we are interested in small
perturbations around $c=-2$. Let us now consider a conformal field theory slightly
deformed from $c=-2$, with $x=1+\epsilon$. Within this generic (non-logarithmic)
CFT, the OPE of a generic operator $\Phi_h$ with itself reads
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h-h_\xi}} \left[\xi(0) +\frac{1}{2} z \partial \xi(0) + \alpha^{(-2)} z^2 L_{-2} \xi(0) + \alpha^{(-1,-1)} z^2 L_{-1}^2 \xi(0) +\dots \right],
\end{equation}
where the $\alpha$ coefficients are fixed by conformal invariance and are
diverging as $\epsilon \rightarrow 0$: $\alpha^{(-2)} = \frac{4h}{27 \epsilon}+\frac{1+2h}{27} + \mathcal{O}(\epsilon)$ and $\alpha^{(-1,-1)} = -\frac{2h}{27 \epsilon}+\frac{4 + h}{27} + \mathcal{O}(\epsilon)$. Note that we showed only the channel
of $\xi$ (with dimension $h_\xi = h_{1,5} = 1 + \frac{3}{2} \epsilon + \mathcal{O}(\epsilon^2)$)
on the right-hand side. Let us
now introduce the field $\phi = (L_{-1}^2 - 2 L_{-2}) \xi$
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h-h_\xi}} \left[\xi(0) +\frac{1}{2} z \partial \xi(0) + \alpha^{(-1,-1)} z^2 \phi(0) + (2 \alpha^{(-1,-1)}+ \alpha^{(-2)})z^2 L_{-2} \xi(0) +\dots \right].
\end{equation}
Remark that we got rid of one of the diverging terms this way as
$\alpha_{\rm reg} = 2 \alpha^{(-1,-1)}+ \alpha^{(-2)} = \frac{9+4h}{27} + \mathcal{O}(\epsilon) $.
At this point, we have no choice but to admit that there exists another field in the
theory with dimension $3$ at $c=-2$ to cancel the last diverging term. As we already
discussed, this field has to be $\Phi_{1,7}$ with conformal weight
$h_{\psi} \equiv h_{1,7} = 3 + 3 \epsilon + \mathcal{O}(\epsilon^2)$. The OPE thus reads
\begin{align}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h-1}} \left[ z^{h_{\xi} - 1} \right. & \xi(0) +\frac{z^{h_{\xi}}}{2} \partial \xi(0) + \alpha^{(-1,-1)} z^{h_{\xi} + 1} \phi(0) \notag \\
& \left. + \alpha_{\rm reg} z^{h_{\xi} + 1} L_{-2} \xi(0) + z^{h_{\psi} - 1} \Phi_{1,7} (0) +\dots \right] .
\end{align}
We define the field $\psi(z)$ as
\begin{equation}
\displaystyle \Phi_{1,7} (z) = \frac{\alpha^{(-1,-1)} \Braket{\phi | \phi}}{\beta(\epsilon)} \psi (z) - \alpha^{(-1,-1)} \phi(z),
\end{equation}
where $\beta = - \frac{ \Braket{\phi | \phi}}{h_{\psi}-h_{\xi}-2}$. We finally
notice that $\alpha^{(-1,-1)} (h_{\psi}-h_{\xi}-2)=-\frac{h}{9}+\mathcal{O}(\epsilon)$,
so that the OPE becomes regular at $c=-2$ and involves logarithms
\begin{equation}
\displaystyle \Phi_h(z) \Phi_h(0) \sim \frac{a_\Phi}{z^{2h-1}} \left[ \xi(0) +\frac{z}{2} \partial \xi(0) + \frac{9+4h}{27} z^{2} L_{-2} \xi(0) + \frac{h}{9} z^2 (\psi (0) + \phi (0) \log z) +\dots \right].
\end{equation}
One can check that the operators $\psi$ and $\phi$ defined this way satisfy the
usual OPEs for logarithmic operators. In particular, it is straightforward to show that
\begin{equation}
\displaystyle \left\langle \phi(z) \psi(0)\right\rangle = \frac{\beta_{1,7}}{z^6},
\end{equation}
where
\begin{equation}
\displaystyle \beta_{1,7} = \lim_{\epsilon \rightarrow 0} \beta (\epsilon) = - \lim_{\epsilon \rightarrow 0} \frac{ \Braket{\phi | \phi}}{h_{\psi}-h_{\xi}-2} = - 18 ,
\end{equation}
as $\Braket{\phi | \phi} = 27 \epsilon + \mathcal{O}(\epsilon^2)$. We find the
same $\beta$ parameter as \cite{KauschGaberdiel} but with a different (technically simpler, but less rigorous) argument
which only involves computation of a few Virasoro commutators.
We now turn to a more general LCFT with central charge $c=1-\frac{6}{x_0 (x_0+1)}$.
Using the previous results, we conjecture that the indecomposability parameter $\beta$
for a generic Jordan cell with structure~\eqref{eq_general_proj}, can be computed from
small deformations around this theory $x=x_0+ \epsilon$ as
\begin{equation}
\label{b_formula}
\displaystyle \boxed{\beta = - \lim_{\epsilon \rightarrow 0} \frac{ \Braket{\phi | \phi}}{h_{\psi}-h_{\xi}-n} = - \frac{\left. \frac{\rm d}{\mathrm{d} \epsilon} \Braket{\phi | \phi}\right|_{\epsilon=0} }{\left. \frac{\rm d}{\mathrm{d} \epsilon} (h_{\psi}-h_{\xi}) \right|_{\epsilon=0}} } ,
\end{equation}
where $n=(\left. h_{\psi}-h_{\xi})\right|_{\epsilon=0}$.
We will show in the
two next sections that eq.~\eqref{b_formula} is consistent with the previous
studies and agrees very well with numerical results. The problem now reduces to identifying Jordan
cells of a given theory. This problem is fairly well understood, and we shall now turn
to concrete lattice examples to illustrate this.
\section{Lattice models and algebraic considerations}
Although the systematic study of Virasoro indecomposable modules can be
analyzed on the Virasoro side in an rather abstract way \cite{KauschGaberdiel,KytolaRidout},
it is also instructive to analyze how indecomposability arises directly from lattice
regularizations \cite{RS3,PRZ}.
The structure of the Virasoro staggered modules can be predicted from the analysis
of the projective modules of associative lattice algebras such as the Temperley-Lieb
algebra \cite{RS3, RP1,RP2}. These results are of course consistent with those of the
Virasoro-based approach. This section follows closely the results of Ref. \cite{RS3}.
\subsection{`Dense' LCFTs from Temperley-Lieb algebra, XXZ spin-1/2 chain, dense loops and supersymmetry}
\begin{figure}
\begin{center}
\begin{pspicture}(0,0)(9,6)
\rput[Bc](1.0,0.5){$\bar{\square} \otimes \square$}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{->}(0,1.0)(1.0,2.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{->}(1.0,2.0)(0.0,3.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{->}(1.0,2.0)(2.0,1.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{->}(2.0,3.0)(1.0,2.0)
\rput[Bc](3.0,2.0){$=$}
\psellipticarc[linecolor=red,linewidth=1.0pt]{->}(6.0,2.0)(0.71,0.71){135}{225}
\psellipticarc[linecolor=red,linewidth=1.0pt]{->}(4.0,2.0)(0.71,0.71){315}{45}
\rput[Bc](5.0,0.5){$p_B$}
\rput[Bc](6.5,2.0){$+$}
\psellipticarc[linecolor=red,linewidth=1.0pt]{<-}(8.0,1.0)(0.71,0.71){45}{135}
\psellipticarc[linecolor=red,linewidth=1.0pt]{<-}(8.0,3.0)(0.71,0.71){225}{315}
\rput[Bc](8.0,0.5){$1-p_B$}
\rput[Bc](1.0,3.5){$\square \otimes \bar{\square}$}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{<-}(0,4.0)(1.0,5.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{<-}(1.0,5.0)(0.0,6.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{<-}(1.0,5.0)(2.0,4.0)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{<-}(2.0,6.0)(1.0,5.0)
\rput[Bc](3.0,5.0){$=$}
\psellipticarc[linecolor=red,linewidth=1.0pt]{<-}(6.0,5.0)(0.71,0.71){135}{225}
\psellipticarc[linecolor=red,linewidth=1.0pt]{<-}(4.0,5.0)(0.71,0.71){315}{45}
\rput[Bc](5.0,3.5){$1-p_A$}
\rput[Bc](6.5,5.0){$+$}
\psellipticarc[linecolor=red,linewidth=1.0pt]{->}(8.0,4.0)(0.71,0.71){45}{135}
\psellipticarc[linecolor=red,linewidth=1.0pt]{->}(8.0,6.0)(0.71,0.71){225}{315}
\rput[Bc](8.0,3.5){$p_A$}
\end{pspicture}
\end{center}
\caption{Graphical representation of the Temperley-Lieb-based vertex model. The
lattice consists of alternating arrows going up for $i$ even and down for $i$ odd,
where $i=1,\dots,L=2N$ corresponds to the horizontal (space) coordinate. The system
has free boundary conditions in the horizontal direction and periodic in the vertical
(imaginary time) direction. We choose each vertex according to its probability; this
draws a dense loop configuration on the lattice. Each closed loop carries a weight $n=q+q^{-1}$.
In the supersymmetric language, the alternating $\square,\bar{\square}$ representations
correspond to a lattice orientation, conserved along each loop. The system is isotropic
when $p_A = p_B$, while the transition occurs when $p_A = 1-p_B$.}
\label{figTL}
\end{figure}
\subsubsection{Temperley-Lieb algebra}
Underlying most of the models we shall consider is the Temperley-Lieb algebra
$TL_{2N}(q)$. The algebra $TL_{2N}(q)$ defined on $L=2N$ ($N \in \mathbb{N}/2$)
strands consists of all the words written with the $2N-1$ generators $e_i$ ($1 \leq i \leq 2N-1$), subject to the relations
\begin{subequations} \label{TLdef}
\begin{eqnarray}
\left[ e_i , e_j \right] &=&0 \ (\left|i-j \right| \geq 2 )\\
e_i ^2 &=& n e_i\\
e_i e_{i \pm 1} e_i &=& e_i
\end{eqnarray}
\end{subequations}
with
\begin{equation}
\displaystyle{ n=q+q^{-1}}.
\end{equation}
We also define $q=\mathrm{e}^{i \gamma}$ and $\gamma = \frac{\pi}{x+1}$.
The action of $e_i$ as an operator acting on $2N$ strands can be naturally
represented graphically. This constructs what we shall refer to as loop or adjoint representation.
The standard modules of the TL algebra are well known, and have dimensions
\begin{equation}
\displaystyle{d_j = \left( \begin{array}{c} 2N \\ N + j \end{array} \right) - \left( \begin{array}{c} 2N \\ N + j +1 \end{array} \right)},
\end{equation}
where in the geometrical language, $2j$ is the number of `through lines' or `strings'
that propagate along the imaginary time direction. We have $j \in \mathbb{N}$, restricted
to the condition $N + j \in \mathbb{N}$, so if $N$ is a half-integer ($L=2N$ odd), so must
be $j$. These standard modules are irreducible for $q$ generic ({\it i.e.} not a root of unity).
We consider a two-dimensional model defined by the transfer matrix
\begin{equation}
\displaystyle{T = \prod_{i=1}^{N-1} \left( p_B + (1-p_B) \ e_{2i} \right) \prod_{i=1}^{N} \left( (1-p_A) + p_A \ e_{2i-1} \right) } ,
\end{equation}
which acts on a given TL module. This definition is valid for $L=2N$ even
but it can be readily adapted to an odd number of sites. We will mainly work
with three different representations: geometrical (adjoint), 6-vertex, and supersymmetric.
Using the geometrical (loop) representation of $TL_{2N}(q)$, we obtain a dense loop model,
where each closed loop carries a weight $n$ (fugacity). A graphical representation of this
vertex model is given Fig.~\ref{figTL}.
It sometimes helps to think in terms of a Q-state Potts model whose high-temperature expansion
consists of drawing dense loop configurations with fugacity $n=\sqrt{Q}$.
In the strong anisotropy limit $p_A \rightarrow 0$ with $p_A / (1-p_B)$ fixed, we can extract
the Hamiltonian of the equivalent one-dimensional quantum system. It reads, up to an irrelevant constant,
\begin{equation}
\displaystyle{ H = - \varepsilon \sum_{i=1}^{N-1} e_{2i} - \varepsilon^{-1} \sum_{i=1}^{N} e_{2i-1}},
\end{equation}
where $\varepsilon = \sqrt{p_A /(1-p_B)}$. The system is isotropic when $p_A = p_B$, while the transition
occurs when $p_A = 1-p_B$. Hereafter, we will always consider the case $\varepsilon = 1$.
Finally, we recall some useful relations for this integrable model. Using Bethe ansatz, one can show
\cite{PottsBethe} that the mean value of the TL generators on the groundstate is
\begin{equation}
\displaystyle e_{\infty} = \sin^2 \gamma \int_{-\infty}^{+\infty} \frac{{\rm d} x}{\cosh \pi x} \ \frac{1}{\cosh 2 \gamma x - 2 \cos \gamma}.
\label{e_inf}
\end{equation}
Meanwhile, the dispersion relation of the quasi-particles is $\varepsilon \sim v_F |k - k_F|$, where the sound velocity reads
\begin{equation}
\displaystyle v_{F} = \frac{\pi \sin \gamma}{\gamma}.
\end{equation}
\subsubsection{6-vertex or XXZ spin chain representation}
Another natural representation of the Temperley-Lieb algebra is provided by the
6-vertex model. We write $n = \mathrm{e}^{i \gamma} + \mathrm{e}^{-i \gamma}$ and
$q = \mathrm{e}^{i \gamma}$. The Hamiltonian limit of the 6-vertex model is the XXZ
chain, with Hilbert space $ \mathcal{H}_{\rm XXZ} = (\mathbb{C}^2)^{\otimes 2N} $. We will
focus on this limit hereafter. The Temperley-Lieb generators in this representation read
\begin{equation}
e_i = \mathbb{I} \otimes \mathbb{I} \otimes \dots \otimes \left( \begin{array}{cccc}
0 & 0 & 0 & 0 \\
0 & q^{-1} & -1 & 0 \\
0 & -1 & q & 0 \\
0 & 0 & 0 & 0 \end{array} \right) \otimes \dots \otimes \mathbb{I},
\label{e_eiXXZ}
\end{equation}
where we have used the basis $\lbrace \Ket{\uparrow \uparrow}, \Ket{\uparrow \downarrow}, \Ket{\downarrow \uparrow}, \Ket{\downarrow \downarrow} \rbrace$ of $ (\mathbb{C}^2)^{\otimes 2}$, the tensor product of the Hilbert spaces
of the sites $i$ and $i+1$. We can check that the generators indeed satisfy the TL algebra,
with $n = q + q^{-1}$. We can express the Hamiltonian of the XXZ chain in terms of the usual Pauli operators
\begin{equation}
\displaystyle H = \frac{1}{2} \sum_{i=1}^{2N-1} \left( \sigma^x_i \sigma^x_{i+1} + \sigma^y_i \sigma^y_{i+1} + \frac{q + q^{-1}}{2} \sigma^z_i \sigma^z_{i+1} \right) + \frac{q - q^{-1}}{4} \left( \sigma^z_1 - \sigma^z_{2N} \right) - N \cos \gamma.
\end{equation}
Note that this XXZ Hamiltonian (or equivalently the transfer matrix of the 6-Vertex model)
commutes \cite{PasquierSaleur} with the generators of the quantum group $U_q(\mathfrak{sl}_2)$.
\subsubsection{Supersymmetric representation (SUSY)}
We introduce in this paragraph a supersymmetric formulation of our model, which provides
another natural representation of $TL_{2N}(q)$ \cite{RS1}. We consider that each edge of
our two-dimensional lattice carries a $\mathbb{Z}_2$ graded vector space of dimension $n+m|m$.
We choose these vector spaces to be the fundamental $\square$ of the Lie superalgebra
$\mathfrak{gl}(n+m|m)$ for $i$ odd (corresponding to down arrows of Fig.~\ref{figTL}),
and the dual $\bar{\square}$ for $i$ even (up arrows). The transfer matrix (or the Hamiltonian)
then acts on the graded tensor product $\mathcal{H} = (\square \otimes \bar{\square})^{\otimes N}$.
The TL generators are simply quadratic Casimir invariants, providing a natural generalization
of the Heisenberg chain to the $\mathfrak{gl}(n+m|m)$ algebra. We can check that a loop expansion
of the transfer matrix yields a dense loop model with a weight $\mathrm{str} \ \mathbb{I} = n + m -m = n$
for each closed loop as expected.
There is a continuum quantum field theory associated with this spin chain, which turns out
\cite{RS1} to be a non-linear $\sigma$-model on complex projective space
$\mathbb{CP}^{n+m-1|m} = \mathrm{U}(m+n|m) / (\mathrm{U}(1) \times \mathrm{U}(m+n-1|m)) $ at
topological angle $\theta = \pi$.
\subsubsection{Continuum limit: the generic case}
Let us also say a few words about the continuum limit. The XXZ chain with
$q=\mathrm{e}^{i \gamma}$, $\gamma = \pi / (x+1)$ is described by a CFT
\cite{PasquierSaleur} with central charge
\begin{equation}
\displaystyle c = 1 - \frac{6}{x (x+1)}.
\end{equation}
Consider the Verma modules $\mathcal{V}_{r,s}$, spanned by the action of
the Virasoro generators $L_n$ with $n<0$ on the highest weight state with
conformal dimension $h_{r,s}$ given by the Kac formula
\begin{equation}
\displaystyle h_{r,s} = \frac{ \left[(x+1)r - xs \right]^2 - 1}{4 x (x+1)}.
\end{equation}
When we take the scaling limit of the XXZ chain, the conformal weights that
occur in the spectrum are $h_{1,1+2j}$ in the Kac table, where $j=n, \ n \in \mathbb{N}$
if L even and $j=n/2, \ n \in 2\mathbb{N}+1$ if L odd. For $q$ generic (not a root of unity),
there is a single null vector in the Verma module $\mathcal{V}_{1,1+2j}$ at conformal weight
$h_{1,-1-2j}$ that must be set to zero in order to obtain a simple (irreducible) module.
Hence, we define the standard (Kac) modules $r_j = r_{1,1+2j} = \mathcal{V}_{1,1+2j} / \mathcal{V}_{1,-1-2j} $
which are irreducible for $q$ generic. The character\footnote{The modular parameter $q$ in
characters and partition functions has of course nothing to do with $q=\mathrm{e}^{i \gamma}$
that parametrizes the weight of closed loops in the Temperley-Lieb algebra. We keep the same
notations and hope this will not confuse the reader.} of the module $r_j$ reads
\begin{equation}
\displaystyle K_j = \mathrm{Tr}_{r_j} \ q^{L_0-c/24} = \frac{q^{h_{1,1+2j}-c/24}-q^{h_{1,-1-2j}-c/24} }{P(q)}.
\end{equation}
where $P(q)$ is the inverse of the Euler partition function, and is related to the Dedekind $\eta$ function
\begin{equation}
\displaystyle P(q) = \prod_{n=1}^{\infty} (1 - q^n) = q^{-1/24} \eta (q).
\end{equation}
The partition function of the sector of spin $S_z$ reads
\begin{equation}
\displaystyle Z^{\mathrm{XXZ}}_{S_z} = \mathrm{Tr} \ q^{L_0-c/24} = \sum_{j=|S_z|}^{\infty} K_j ,
\end{equation}
the global partition function of the $U_q(\mathfrak{sl}_2)$ XXZ spin chain is then readily obtained
\begin{equation}
\displaystyle Z^{\mathrm{XXZ}} = \sum_{S_z=-\infty}^{+\infty} Z^{\mathrm{XXZ}}_{S_z} = \sum_{j=0}^{\infty} (2 j + 1) K_j .
\end{equation}
The partition function of the superspin chain with $\mathfrak{gl}(n+m|n)$ symmetry is given
by a similar expression where one replaces $2j+1$ by the dimensions of the irreducible
representations of the commutant $\mathcal{A}_{n+m|n}(2N)$ of $TL_{2N}(q)$ in the SUSY
representation \cite{RS2}. The algebra $\mathcal{A}_{n+m|n}$ is in fact much larger than $\mathfrak{gl}(n+m|n)$.
These results were obtained for $q$ generic but are supposed to remain correct even
when $q$ is a root of unity even though the Virasoro standard modules are no longer
irreducible in general.
\subsection{`Dilute' LCFTs from the integrable $O(n)$ model}
We will also consider a fundamentally different version of the previous LCFTs using
the integrable dilute $O(n)$ on the square lattice. We shall refer to these theories
as `dilute' as opposed to the `dense' ones based on the Temperley-Lieb algebra. This
denomination obviously refers to the dense or dilute nature of the underlying loop gas.
Therefore, we describe in this section the $O(n)$ model defined on an annulus of
width $2N$. It corresponds to a dilute loop model where closed loops carry a weight
$n$; we shall focus here only on the dilute phase. This model also possesses a dense phase
which is in the same universality class as the dense loop model. Note that the case $n \rightarrow 0$
is obviously relevant for the physics of polymers. In terms of spin chains, it is described
by a $S=1$ $U_q(\mathfrak{sl}_2)$-invariant chain where the states $S_z=\pm 1$ are viewed as occupied
by parts of loops and $S_z= 0$ as empty. This model also corresponds to $\mathfrak{osp}(n+2m|2m)$
(super)spin chains and to non-linear sigma models with supersphere target space $S^{2m+n-1|2m} \simeq \mathrm{OSp}(2m+n|2m)/\mathrm{OSp}(2m+n-1|2m)$ \cite{RS1}. There is a dilute version of the Temperley-Lieb algebra behind
all these models. We will not go into the details of these different formulations here
but only describe the geometrical setup that will allow us to measure indecomposability
parameters in the next section.
\subsubsection{Lattice model}
\begin{figure}
\begin{center}
\begin{pspicture}(-0.5,-0.5)(14,5.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(-0.5,3)(0.5,4)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(0.5,4)(1.5,3)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(-0.5,3)(0.5,2)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(0.5,2)(1.5,3)
\rput[Bc](0.5,1.5){$\rho_1$}
\psellipticarc[linecolor=red,linewidth=2.0pt]{-}(2,4.5)(0.71,0.71){315}{45}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(2,4.5)(3,5.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(3,5.5)(4,4.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(2,4.5)(3,3.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(3,3.5)(4,4.5)
\rput[Bc](3.0,3.0){$\rho_2$}
\psellipticarc[linecolor=red,linewidth=2.0pt]{-}(4,1.5)(0.71,0.71){135}{225}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(2,1.5)(3,2.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(3,2.5)(4,1.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(2,1.5)(3,0.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(3,0.5)(4,1.5)
\rput[Bc](3.0,0.0){$\rho_3$}
\psellipticarc[linecolor=red,linewidth=2pt]{-}(5.5,2.5)(0.71,0.71){225}{315}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(4.5,4.5)(5.5,5.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(5.5,5.5)(6.5,4.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(4.5,4.5)(5.5,3.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(5.5,3.5)(6.5,4.5)
\rput[Bc](5.5,3.0){$\rho_4$}
\psellipticarc[linecolor=red,linewidth=2pt]{-}(5.5,3.5)(0.71,0.71){45}{135}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(4.5,1.5)(5.5,2.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(5.5,2.5)(6.5,1.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(4.5,1.5)(5.5,0.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(5.5,0.5)(6.5,1.5)
\rput[Bc](5.5,0.0){$\rho_5$}
\psline[linecolor=red,linewidth=2pt]{-}(7.5,4)(8.5,5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(7,4.5)(8,5.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(8,5.5)(9,4.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(7,4.5)(8,3.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(8,3.5)(9,4.5)
\rput[Bc](8,3.0){$\rho_6$}
\psline[linecolor=red,linewidth=2pt]{-}(7.5,2)(8.5,1)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(7,1.5)(8,2.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(8,2.5)(9,1.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(7,1.5)(8,0.5)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(8,0.5)(9,1.5)
\rput[Bc](8,0.0){$\rho_7$}
\psellipticarc[linecolor=red,linewidth=2.0pt]{-}(11.5,3.0)(0.71,0.71){135}{225}
\psellipticarc[linecolor=red,linewidth=2.0pt]{-}(9.5,3.0)(0.71,0.71){315}{45}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(9.5,3)(10.5,4)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(10.5,4)(11.5,3)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(9.5,3)(10.5,2)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(10.5,2)(11.5,3)
\rput[Bc](10.5,1.5){$\rho_8$}
\psellipticarc[linecolor=red,linewidth=2pt]{-}(13.0,2.0)(0.71,0.71){45}{135}
\psellipticarc[linecolor=red,linewidth=2pt]{-}(13.0,4.0)(0.71,0.71){225}{315}
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(12,3)(13,4)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(13,4)(14,3)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(12,3)(13,2)
\psline[linecolor=black,linewidth=0.5pt,arrowsize=5pt]{-}(13,2)(14,3)
\rput[Bc](13,1.5){$\rho_9$}
\end{pspicture}
\end{center}
\caption{Plaquettes and integrable weights for the dilute $O(n)$ model on the square lattice}
\label{figOn}
\end{figure}
Our starting point is the expression of the integrable version of the
dilute $O(n)$ model defined on the square lattice. The $\v{R}$ matrix is the
sum of 9 plaquettes represented graphically in Figure \ref{figOn}
\begin{equation}
\displaystyle \v{R}_j(u) = \sum_{i=1}^{9} \rho_i(u) \mathcal{O}_j^i,
\end{equation}
where $j \in \{1, \dots, 2N-1 \}$ labels the sites.
It satisfies the Yang-Baxter equations for the corresponding integrable weights \cite{Onweights}
\begin{subequations}
\begin{eqnarray}
\rho_1 &=& 1+\frac{\sin u \sin (3 \lambda - u)}{\sin 2 \lambda \sin 3 \lambda} \\
\rho_2 &=& \rho_3 = \frac{\sin (3 \lambda - u)}{\sin 3 \lambda}\\
\rho_4 &=& \rho_5 = \frac{\sin u}{\sin 3 \lambda}\\
\rho_6 &=& \rho_7 = \frac{\sin u \sin(3 \lambda -u)}{\sin 2 \lambda \sin 3 \lambda}\\
\rho_8 &=& \frac{\sin (2 \lambda -u) \sin(3 \lambda -u)}{\sin 2 \lambda \sin 3 \lambda}\\
\rho_9 &=& -\frac{\sin u \sin(\lambda -u)}{\sin 2 \lambda \sin 3 \lambda}
\end{eqnarray}
\end{subequations}
where $n = - 2 \cos 4 \lambda$ is the weight given to every closed loop. Note that
\begin{equation}
\displaystyle \v{R}_i(u=0) = \mathcal{O}_i^1 + \mathcal{O}_i^2 + \mathcal{O}_i^3 + \mathcal{O}_i^8 = \mathrm{Id}.
\end{equation}
One can extract the corresponding 1D Hamiltonian $H= - \left. \frac{\mathrm{d}\v{R}}{\mathrm{d}u}\right|_{u=0}$ using an expansion in the
spectral parameter $u$. We find that the interaction between the sites $i$ and $i+1$ reads
\begin{align}
\displaystyle h_{i,i+1} = \frac{1}{\sin 2 \lambda} \mathcal{O}_i^1 & - \mathrm{cotan} 3 \lambda (\mathcal{O}_i^2 + \mathcal{O}_i^3) + \frac{1}{ \sin 3 \lambda} (\mathcal{O}_i^4 + \mathcal{O}_i^5) \notag \\
+& \frac{1}{ \sin 2 \lambda} (\mathcal{O}_i^6 + \mathcal{O}_i^7) - (\mathrm{cotan} 2 \lambda + \mathrm{cotan} 3 \lambda) \mathcal{O}_i^8 - \frac{\sin \lambda}{\sin 2 \lambda \sin 3 \lambda} \mathcal{O}_i^9.
\label{hamOn}
\end{align}
In order to obtain an integrable Hamiltonian in the case of open (free) boundary
conditions, we would need to consider K matrices at the edges \cite{Skyanin}. Instead,
we here consider the simpler (non-integrable) Hamiltonian $H = -\sum h_{i,i+1}$. In what follows, every
reference to the $O(n)$ model corresponds to this Hamiltonian.
The Fermi velocity and the mean value of $h_{i,i+1}$ on the ground state are \cite{OnBethe}
\begin{equation}
\displaystyle v_F = \frac{\pi}{3 \lambda},
\end{equation}
and
\begin{equation}
\displaystyle h_{\infty} = 2 \int_{-\infty}^{+\infty} {\rm d} k \ \frac{\tanh 3 \lambda k \cosh (5 k \lambda - k \pi) \cosh k \lambda}{\sinh k \pi}.
\end{equation}
\subsubsection{Continuum limit}
The CFT describing the scaling limit of dilute loops has central charge \cite{Onc}
\begin{equation}
\displaystyle c = 1-3 \frac{(4 \lambda - \pi)^2}{2 \lambda \pi},
\end{equation}
with $n= - 2 \cos 4 \lambda$ and $\lambda \in [\pi/4,\pi/2]$. The conformal weights
appearing in the spectrum lie now in the first row of the Kac table $h_{1+2j,1}$.
The trace over the standard module with $2j$ through lines reads
\begin{equation}
\displaystyle K_{j} = \mathrm{Tr}_{r_j} \ q^{L_0-c/24} = \frac{q^{h_{1+2j,1}-c/24}-q^{h_{-1-2j,1}-c/24} }{P (q)}.
\end{equation}
As in the dense case, the partition function of the spin-1 XXZ chain is just the sum
of these characters with a degeneracy $2j+1$. In the case of $\mathfrak{osp}(n+2m|2m)$-invariant
superspin chains, the degeneracy $2j+1$ is replaced by the dimension of the corresponding irreducible
representation of the commutant $\mathcal{B}_{n+2m|2m}(2N)$ of the dilute Temperley-Lieb algebra \cite{RS2}.
\subsection{Indecomposability and lattice Jordan cells at $q$ root of unity on a concrete example: the logarithmic Ising model $\mathcal{LM}(3,4)$}
\label{subsec_ising}
We now turn to the study of indecomposability when $q$ is a root of unity. In this case,
the standard modules of the Temperley-Lieb algebra still exist but may no longer be irreducible.
Read and Saleur \cite{RS3} showed in details that the structure of the projective modules of
the Temperley-Lieb algebra parallels that of several staggered Virasoro modules in the continuum
limit. The general algebraic structure of the XXZ spin chain was illustrated in terms of
`staircase diagrams' as a representation (bimodule) of $TL_{2N}(q) \otimes U_q(\mathfrak{sl}_2)$, or
$TL_{2N}(q) \otimes \mathcal{A}_{n+m|n}(2N)$ in the SUSY case.
In this paper, we will pass directly to the continuum theory and only mention the important
algebraic lattice results when necessary. We refer the interested reader to Ref.~\cite{RS3} for more details.
Let us consider the case $q=\mathrm{e}^{i \pi/4}$, which corresponds to a dense CFT with
central charge $c=\frac{1}{2}$. Loop expansion of the partition function yields a loop
model with fugacity $n=\sqrt{2}$; this is of course the celebrated (logarithmic) Ising model $\mathcal{LM}(3,4)$ \cite{PRZ}.
The Kac formula at $x=3$ reads
\begin{equation}
\displaystyle h_{r,s} = \frac{ \left[4r - 3 s \right]^2 - 1}{48},
\end{equation}
and the values appearing in the spectrum of the $U_q(\mathfrak{sl}_2)$-invariant XXZ
spin chain at $q=\mathrm{e}^{i \pi/4}$ are
\begin{equation}
\displaystyle h_{1,1+2j} = \frac{j(3j-1)}{4}.
\end{equation}
We will focus the continuum limit of a spin chain with an even number of sites so
that $j \in \mathbb{N}$; the case $j = n/2, \ n \in 2\mathbb{N}+1$ is treated in
a similar way. Note that we have $h_{r,s}=h_{-r,-s}=h_{r+3,s+4}$. The character
of the standard module $r_j$ reads
\begin{equation}
\displaystyle K_{j} = \mathrm{Tr}_{r_j} \ q^{L_0-c/24} = \frac{q^{(1- 6 j)^2/48}-q^{(7+6j)^2/48} }{\eta (q)},
\end{equation}
whereas the characters \cite{Characters} of the simple (irreducible) Virasoro module are
\begin{subequations}
\begin{eqnarray}
\chi_{j=4p}&=&
\sum_{n\notin [-2p,-1]}
\frac{q^{(24n-1+24p)^{2}/48}-q^{(24n+17+24p)^{2}/48}}
{\eta(q)},
\\
\chi_{j=4p+1}&=&
\sum_{n\notin [-2p,-1]}
\frac{q^{(24n+5+24p)^{2}/48}-q^{(24n+11+24p)^{2}/48}}
{\eta(q)},
\\
\chi_{j=4p+2}&=&
\sum_{n\notin [-2p-1,-1]}
\frac{q^{(24n+11+24p)^{2}/48}-q^{(24n+29+24p)^{2}/48}}
{\eta(q)},
\\
\chi_{j=4p+3}&=&
\sum_{n\notin [-2p-1,-1]}
\frac{q^{(24n+17+24p)^{2}/48}-q^{(24n+23+24p)^{2}/48}}
{\eta(q)}.
\end{eqnarray}
\end{subequations}
We can see using these character identities that the standard
modules $r_j$, $j \in \mathbb{N}$ are no longer irreducible.
The decomposition onto the simple characters indeed yields
\begin{subequations}
\begin{eqnarray}
K_{j=4p}&=&\chi_{4p}+\chi_{4p+3}, \\
K_{j=4p+1}&=&\chi_{4p+1}+\chi_{4p+2}, \\
K_{j=4p+2}&=&\chi_{4p+2}+\chi_{4p+5}, \\
K_{j=4p+3}&=&\chi_{4p+3}+\chi_{4p+4}.
\end{eqnarray}
\end{subequations}
Let $R_{j}$ be the simple Virasoro module with conformal weight
$h_{1,1+2j}$ and character $\chi_j$.
From lattice algebraic considerations, we know that the standard
modules $r_j$ must thus have the structure
\begin{equation}
r_{j}=~~~~~\left\{\begin{array}{cl}
\begin{array}{ccc}
R_{j}&&\\
&\hskip-.2cm\searrow&\\
&&\hskip-.3cmR_{j+3}
\end{array} &\hbox{$j\equiv0 \ ($mod$ 4)$ or $j\equiv2 \ ($mod$ 4)$}\nonumber\\
\begin{array}{ccc}
R_{j}&&\\
&\hskip-.2cm\searrow&\\
&&\hskip-.3cmR_{j+1}
\end{array} &\hbox{$j\equiv1 \ ($mod$ 4)$ or $j\equiv3 \ ($mod$ 4)$}\nonumber\\
\end{array}\right.
\end{equation}
The arrows represent again the action of the Virasoro algebra.
The bottom submodule is invariant while it is possible to go from
the top to the bottom acting by some element of Virasoro. The top
simple quotient must have a smaller $j$ number than the bottom \cite{RS3}.
Using the standard modules as elementary bricks, it becomes quite
easy to construct the staggered modules of the theory using the knowledge, again from the lattice,
that they must be diamond-shaped. For instance for $j=2$, we expect
\begin{equation}
\mathcal{P}_2=
\begin{array}{ccccc}
&&\hskip-.7cm R_2&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_1&&&\hskip-.3cm R_5 \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_2&&
\end{array}
\label{eq_projisingex}
\end{equation}
with character $P_2=2\chi_{2}+\chi_{1}+\chi_{5}$. This is the only
gluing of standard modules that respects the fact that the conformal
weights must be increasing to the right. For a given theory,
there is always a unique way
to construct a given diamond-shaped module in terms of simple modules\footnote{To be more precise, we mean here that
for a given module $P_j$ of a given theory (dense or dilute),
the values of the subscripts j of the four simples $R_j$ in the diamond $P_j$ are uniquely fixed.}
using only the structure the standard modules.
Using this method, it becomes quite straightforward to guess the general
structure of the Virasoro staggered modules appearing in the theory
\begin{equation}\label{eq_proj_ising}
{\cal P}_{j}=~~~~~\left\{\begin{array}{cl}
\begin{array}{ccc}
R_0&&\\
&\hskip-.2cm\searrow&\\
&&\hskip-.3cm R_3\end{array}&\hbox{$j=0$,}\\
\begin{array}{ccc}
R_1&&\\
&\hskip-.2cm\searrow&\\
&&\hskip-.3cm R_2\end{array}&\hbox{$j=1$,}\\
\begin{array}{ccccc}
&&\hskip-.7cm R_j&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-1}&&&\hskip-.3cm R_{j+3}\\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j&&
\end{array}&\hbox{$j\equiv0$ (mod 4) and $j>0$, or $j\equiv2$ (mod 4),}\\
&\\
\begin{array}{ccccc}
&&\hskip-.7cm R_j&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-3}&&&\hskip-.3cm R_{j+1}\\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j&&
\end{array}&\hbox{$j\equiv1$ (mod 4) and $j>1$, or $j\equiv3$ (mod 4).}\end{array}\right.
\end{equation}
Of course, it is also possible using Temperley-Lieb representation
theory to check that a similar pattern arises from the lattice. Note
also that it may happen that the standard module $r_j$ remains irreducible
for $q$ a root of unity. For example, in the Ising model with $L$ odd,
one finds that $K_{j=4p+3/2}=\chi_{4p+3/2}$ so that what we shall call staggered module
in this case is just the simple module itself ${\cal P}_{j=4p+3/2}=r_{j=4p+3/2}=R_{j=4p+3/2}$.
Let us summarize what we have learned so far concerning the structure of
the scaling limit of the XXZ spin chain at $q=\mathrm{e}^{i \pi/4}$ with
an even number of sites. All the fields of the theory are organized into
staggered modules given by~\eqref{eq_proj_ising}, $j \in \mathbb{N}$.
Recall that each simple module $R_j$ corresponds to a field with conformal
weight $h_{1,1+2j}$. There is a Jordan cell in $L_0$ for every staggered module
with a diamond shape. For such generic staggered module, we note the basis fields
\begin{equation}\label{eq_proj_j}
\mathcal{P}_j=
\begin{array}{ccccc}
&&\hskip-.7cm R_j &&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j_1} &&&\hskip-.3cm R_{j_2} \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j &&
\end{array}
=
\begin{array}{ccccc}
&&\hskip-.7cm \psi^{(j)} &&\\
&\hskip-.2cm\swarrow&\searrow&\\
\xi^{(j)} &&&\hskip-.3cm \rho^{(j)} \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm \phi^{(j)} &&
\end{array}
,
\end{equation}
with $j_1 \leq j \leq j_2$ so that $h_{1,1+2 j_1} \leq h_{1,1+2 j} \leq h_{1,1+2 j_2}$.
We define the logarithmic coupling $\beta_{1,1+2j}$ of this Jordan cell as
\begin{equation}
\displaystyle \beta_{1,1+2j} = \Braket{\phi^{(j)} | \psi^{(j)}}.
\end{equation}
Note that we use the same Kac labels for $\beta$ as those of the conformal weight of $\psi^{(j)}$.
We are now ready to compute this indecomposability parameter using eq.~\eqref{b_formula}
along with $h_\xi = h_{1,1+2 j_1}$ and $h_\psi = h_{1,1+2 j_2}$.
Let us again illustrate this on a concrete example. The first Jordan cell
occurring in the spectrum has $j=2$ and its structure is described by
eq.~\eqref{eq_projisingex}. Defining $x=3+\epsilon$, it is straightforward
to compute $h_{1,3}= \frac{1}{2}+\frac{\epsilon}{8} + \mathcal{O}(\epsilon^2)$ ($j=1$)
and $h_{1,5}= \frac{5}{2}+\frac{3\epsilon}{8}+ \mathcal{O}(\epsilon^2)$ ($j=2$).
At $\epsilon = 0$, the $L_0$ operator expressed in the basis ($\phi^{(2)}$,$\psi^{(2)}$) reads
\begin{equation}
L_0 =
\left( \begin{array}{cc}
5/2 & 1 \\
0 & 5/2 \end{array} \right).
\end{equation}
We remark that we can readily find the relation between $\phi^{(2)}$ and
$\xi^{(2)}$ (up to an irrelevant global normalization factor) because we
know that $\phi^{(2)}$ must be a null-vector
\begin{equation}
\displaystyle \phi^{(2)}(z) = (L_{-2}- \frac{3}{4} L^2_{-1})\xi^{(2)}(z).
\end{equation}
In general, the operator $A$ is either known from general formulas or
computed numerically using the Virasoro algebra.
Straightforward commutations of the $L_{n}$'s modes yield the final result
\begin{equation}
\label{eq_b_ising}
\displaystyle \beta_{1,5} = - \lim_{\epsilon \rightarrow 0} \frac{ \Braket{\phi^{(2)} | \phi^{(2)}} }{h_{1,5}-h_{1,3}-2} = - \dfrac{35}{24} .
\end{equation}
\subsection{General structure of the staggered Virasoro modules}
\label{generalstructure}
It should be clear that the path followed in the previous subsection
can be extended to all our models. In this paper, we focus on the dense
and dilute versions of the minimal logarithmic models $\mathcal{LM}(1,p)$
and $\mathcal{LM}(p,p+1)$, $p\in \mathbb{N}$. The former choice corresponds
to $x=1/p$ and the latter to $x=p$, with $q=\mathrm{e}^{i \pi /(x+1)}$. The
structure of the standard modules in all cases can be inferred from characters
identities, and the staggered modules can then be built using the standards as elementary
bricks. Using character identities, it is possible to convince oneself that the
two following statements should hold
\begin{itemize}
\item the $\mathcal{LM}(1,p)$ theories ($x=1/p$) have the same
structure as their `dual' $\mathcal{LM}(p,p+1)$ theories with $x=p$,
\item a `dilute' LCFT based on the $O(n)$ model has the same
structure as the `dense' LCFT with the same loop fugacity $n=q+q^{-1}$.
\end{itemize}
We say that two theories have the same structure when the
$\mathcal{P}_j$ modules in both theories have the same expression in
terms of the simple modules $R_j$, that is to say that they are characterized by the same values of $j_1$ and $j_2$ in eq.~\eqref{eq_proj_j}.
This does not mean that the modules are the same, in particular, they may be characterized by different
indecomposability parameters, and the simple modules $R_j$ are completely different as they are over different algebras.
For example, the staggered modules in the theory with $x=1/3$ have the same structure in terms of simple modules as those of the Ising model,
given by eq.~\eqref{eq_proj_ising}.
Therefore, everything boils down to the study of the staggered modules\footnote{Note that what we call staggered modules here are nothing but the `scaling limit' of the Temperley-Lieb projective modules that arise in lattice models. They may or may not be indecomposable diamonds depending on the value of the `spin' $j$.}
of $\mathcal{LM}(p,p+1)$ theories with $x=p$.
For such theories, we find that the staggered modules have the following form
\begin{equation}
{\cal P}_{j}=~~~~~\left\{\begin{array}{cl}
\begin{array}{ccccc}
&&\hskip-.7cm R_j&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-1}&&&\hskip-.3cm R_{j+p}\\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j&&
\end{array}&\hbox{$j\equiv0$ \ (mod $\frac{p+1}{2}$) and $j>0$,}\nonumber\\
&\nonumber\\
\begin{array}{ccccc}
&&\hskip-.7cm R_j&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-2}&&&\hskip-.3cm R_{j+p-1}\\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j&&
\end{array}&\hbox{$j\equiv\frac{1}{2}$ \ (mod $\frac{p+1}{2}$) and $j>1$,}\nonumber\\
&\nonumber\\
\vdots \\
\begin{array}{ccccc}
&&\hskip-.7cm R_j&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-p}&&&\hskip-.3cm R_{j+1}\\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_j&&
\end{array}&\hbox{$j\equiv\frac{p-1}{2}$ \ (mod $\frac{p+1}{2}$) and $j>p-1$,}\nonumber\\
&\nonumber\\
R_j &\hbox{$j\equiv\frac{p}{2}$ \ (mod $\frac{p+1}{2}$).}\nonumber\\
&\nonumber\\
\end{array}\right.
\end{equation}
Once again, each diamond module corresponds to a Jordan cell for $L_0$ involving the null-field $\phi_j(z)$, with
logarithmic coupling $\beta_{1,1+2j}$ ({\it resp.} $\beta_{1+2j,1}$) in the
dense ({\it resp.} dilute) case given by eq.~\eqref{b_formula}. For $j<p-1$,
only the top standard module of the diamond remains in the staggered.
\section{Numerical measure of indecomposability parameters from lattice models}
\label{numericalmethod}
While the analysis of symmetries of the lattice models provides information about the general structure of the Virasoro indecomposable modules,
getting more detailed information about the action of the Virasoro generators in these modules---such as the numerical values of the
indecomposability parameters---is more challenging.
There are many difficulties to overcome in this kind of analysis. One of the most important ones concerns
the proper normalization of the state $\phi$, which obeys $\Braket{\phi | \phi} = 0$.
For $c=0$ and the Jordan cell of the energy-momentum tensor, Ref.~\cite{DJS} used a trick
that led to the determination of
$b=-5/8$ for percolation and $b=5/6$ for self-avoiding random walks (dilute polymers).
We follow here a different route, and propose a more general method which allows us to study
other Jordan cells with an $A$ operator more complicated than $L_{-2}$.
The method turns out te be quite accurate so that indecomposability parameters can be
determined numerically, with almost the same precision as for critical exponents.
\subsection{Jordan cells and lattice scalar product}
\subsubsection{Spectrum and Jordan cells}
We consider a generic Jordan cell at level $h$ in $L_0$. As before, we normalize
our states such that in the basis ($\phi$,$\psi = A \xi$), $L_0$ reads
\begin{equation}
L_0 =
\left( \begin{array}{cc}
h & 1 \\
0 & h \end{array} \right).
\end{equation}
It is well known that the operator $L_0$ can be related to the scaling limit
of a Hamiltonian at a critical point. For a system of length $L=2N$, we have
\begin{equation}
\displaystyle H = E_S + E^{\infty}_0 L + \frac{\pi v_F}{L} \left( L_0 - \frac{c}{24}\right) + \mathcal{O}(L^{-2}),
\end{equation}
where $E_S$ is a (non-universal) surface energy and $E_0^{\infty}$ is the bulk
energy density. Hence, we expect non-diagonalizability also in our lattice
Hamiltonians (or transfer matrices) that mimic the continuum limit behavior.
The central charge and the critical exponents of a given model can be readily
measured numerically using finite-size corrections in $L=2N$ to the
eigenvalues of $H$. If we note $E_0$ the energy of the fundamental and $E_{\phi}$ that
of a given excitation, we have the well-known relations \cite{CardyScaling}
\begin{subequations}
\begin{eqnarray}
\displaystyle E_0(L) &=& E_0(\infty)L + E_0^S -\frac{v_F \pi c}{24 L} + \dots \\
\displaystyle E_{\phi}(L) - E_{0}(L) &=& E_{\phi}^S- E_0^S +\frac{v_F \pi}{L} h_{\phi} + \dots
\end{eqnarray}
\end{subequations}
with most of the times $E_{\phi}^S= E_0^S$. We here assumed that all the
conformal dimensions $h_{\phi}$ were positive; if that is not the case,
one introduces the usual concept of effective central charge $c_{\mathrm{eff}}=c-24 h_{\rm min}$.
Using these formulae, it is then a simple matter to identify the different
eigenstates of the Hamiltonian which corresponds to $ \Ket{\psi}$, $\Ket{\phi}$
or $\Ket{\xi}$ in the continuum limit. To do so, we use the Arnoldi algorithm \cite{Arnoldi} to
get the eigenvalues and the corresponding Schur vectors for the lowest excitations
of the spectrum. We then apply some variant of the Gauss-Jordan algorithm to put
the reduced Schur (upper triangular) matrix into Jordan canonical form\footnote{This step involves choosing
a rule to declare that two numbers $\lambda_1$ and $\lambda_2$ are equal if and only if $|\lambda_1-\lambda_2|<\varepsilon$
with typically $\varepsilon \sim 10^{-8}$.}.
Let us suppose that we identify a Jordan cell in the Hamiltonian which
corresponds to an energy $E(L) $.
We normalize the states to prepare the comparison with CFT. In the
basis $\lbrace \Ket{\phi^{(L)}}, \Ket{\psi^{(L)}} \rbrace$, the Hamiltonian
for a system with $L=2N$ sites reads
\begin{equation}
\displaystyle H^{(L)} - E_0(L) \mathrm{Id}= \frac{\pi v_F}{L} \left( \begin{array}{cc} h^{(L)} & 1 \\ 0 & h^{(L)} \end{array} \right),
\label{eq_latticeH}
\end{equation}
where $v_F$ is the Fermi velocity and $h^{(L)} = \frac{L}{\pi v_F} (E(L) - E_0(L)) $.
Note that we have $\lim_{L \to \infty} h^{(L)} = h$, so the matrix expression in eq.~
\eqref{eq_latticeH} goes to $L_{0}$ in the continuum limit.
\subsubsection{Lattice scalar products}
In order to measure $\beta = \Braket{\psi | \phi}$, we first need to define a
`scalar product' that goes to the Virasoro bilinear form in the scaling limit.
The construction of lattice `scalar products' which go to the Virasoro form
in the continuum has already been studied in great details in Ref. \cite{DJS}.
All scalar products must be chosen such that $L_0^{\dag} = L_0$ in the
underlying CFT. This means that we want the Hamiltonian $H$ to be hermitian
for these scalar products. Of course, as we deal with non-unitary theories,
there may be negative-norm states so what we call here scalar product is nothing
but a sesquilinear form. Obviously, the non-hermitianity of $H$ for the usual scalar
product was the reason for its non-diagonalizability in the first place. Let us recall
the expression of the scalar product for the different representations of the (dense/dilute)
Temperley-Lieb algebra:
\paragraph{XXZ:} The scalar product is the usual bilinear form on $\mathbb{C}$ without
complex conjugation, that is, treating $q$ as a formal parameter. For example, on $L=4$ sites,
the vector $\Ket{\phi} = \Ket{\uparrow \uparrow \downarrow \downarrow} + q \Ket{\uparrow \uparrow \uparrow \uparrow} $
has norm $\Braket{\phi|\phi} = 1+q^2$. Note that if we had considered the usual scalar product
on $\mathbb{C}$, we would have found $1+|q|^2$ instead.
\paragraph{LOOP:} The correct scalar product is obtained gluing the mirror image of the first
state on top of the second one. Each closed loop carries a weight $n=q+q^{-1}$. This is of
course the usual form used in Temperley-Lieb representation theory. For instance, the scalar
product between the two states
$\Ket{\alpha} = \Ket{
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(3,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,1.0)(1.5,1.42){180}{360}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,1.0)(0.5,0.71){180}{360}
\end{pspicture}
}$
and
$\Ket{\beta} = \Ket{
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(3,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,1.0)(0.5,0.71){180}{360}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2.5,1.0)(0.5,0.71){180}{360}
\end{pspicture}
}$
is
$\Braket{\alpha|\beta} = \Braket{
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(3,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,0.0)(1.5,1.42){0}{180}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,0.0)(0.5,0.71){0}{180}
\end{pspicture}
|
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(3,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,1.0)(0.5,0.71){180}{360}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2.5,1.0)(0.5,0.71){180}{360}
\end{pspicture}
}=
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(3,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,0.25)(1.5,1.42){0}{180}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,0.25)(0.5,0.71){0}{180}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,0.25)(0.5,0.71){180}{360}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2.5,0.25)(0.5,0.71){180}{360}
\end{pspicture}
=n$.
The case with a non-zero number of strings $2j$ is treated in a similar fashion.
Finally, in the case of the dilute $O(n)$ model, the scalar product between two states
is chosen to be zero if the empty sites (marked as dots) are not the same. For example for $L=6$, $ \Braket{
\psset{xunit=2mm,yunit=2mm}
\psset{dotsize=2pt 0}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2,1.0)(2,1.42){180}{360}
\psdot(2,1.0)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2,1.0)(1,0.71){180}{360}
\psdot(5,1.0)
\end{pspicture}
|
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,1.0)(0.5,0.71){180}{360}
\psdot(2,1.0)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(3.5,1.0)(0.5,0.71){180}{360}
\psdot(5,1.0)
\end{pspicture}
}
=
\Braket{
\psset{xunit=2mm,yunit=2mm}
\psset{dotsize=2pt 0}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2,0.0)(2,1.42){0}{180}
\psdot(2,0.0)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(2,0.0)(1,0.71){0}{180}
\psdot(5,0.0)
\end{pspicture}
|
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,1.0)(0.5,0.71){180}{360}
\psdot(2,1.0)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(3.5,1.0)(0.5,0.71){180}{360}
\psdot(5,1.0)
\end{pspicture}
}=n$,
whereas
$
\Braket{
\psset{xunit=2mm,yunit=2mm}
\psset{dotsize=2pt 0}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,1.0)(1.5,1.42){180}{360}
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(1.5,1.0)(0.5,0.71){180}{360}
\psdot(4,1.0)
\psdot(5,1.0)
\end{pspicture}
|
\psset{xunit=2mm,yunit=2mm}
\begin{pspicture}(0,0)(5.5,1)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(0.5,1.0)(0.5,0.71){180}{360}
\psdot(2,1.0)
\psellipticarc[linecolor=black,linewidth=1.0pt]{-}(3.5,1.0)(0.5,0.71){180}{360}
\psdot(5,1.0)
\end{pspicture}
}=0$.
\paragraph{SUSY:} We use the usual scalar product in Fock space. There are negative
norm states because of the use of the dual representation. For example, let us consider
the $\mathfrak{sl}(2|1)$ case still on $L=4$ sites. A precise definition of this chain
will be given in the following. The important point here is that each site must be occupied
by one particle which can be either a fermion $\{ f_i, f_j^\dag \} = (-1)^{i+1} \delta_{ij}$,
or a Schwinger boson $[ b_{i,\sigma}, b^\dag_{j,\sigma'} ] = \delta_{ij} \delta_{\sigma\sigma'}$,
with $\sigma \in \{ \uparrow, \downarrow \}$. Let us consider the state $\Ket{\phi} = b^\dag_{1 \uparrow} f^\dag_{2} b^\dag_{3 \downarrow} b^\dag_{4 \uparrow}\Ket{0}$. Its norm is $\Braket{\phi|\phi}= \Braket{0 \right| b_{4 \uparrow} b_{3 \downarrow} f_{2} b_{1 \uparrow} b^\dag_{1 \uparrow} f^\dag_{2} b^\dag_{3 \downarrow} b^\dag_{4 \uparrow} \left|0} = -1$ because of the
fermionic operator $f^\dag_{2}$ of the dual representation which satisfies $\{f_{2},f^\dag_{2} \}=-1$.
\subsection{Virasoro algebra regularization on the lattice}
There is a last difficulty that one must tackle in order to define a proper version
of $\beta$ on the lattice. Remark that the Jordan cell in~\eqref{eq_latticeH} is invariant
under a global rescaling of the basis states $\Ket{\phi^{(L)}} \rightarrow \alpha \Ket{\phi^{(L)}} $
and $\Ket{\psi^{(L)}} \rightarrow \alpha \Ket{\psi^{(L)}} $. Unfortunately, the scalar product
between the states is not invariant under such transformation $ \Braket{\psi^{(L)}|\phi^{(L)}} \rightarrow |\alpha|^2 \Braket{\psi^{(L)}|\phi^{(L)}}$. Hence, we need to normalize the state $\Ket{\phi^{(L)}} $ so that it goes
precisely to $\Ket{\phi}= A \Ket{\xi}$ in the continuum limit.
Let $\Ket{\phi^{(L)}} = \alpha \Ket{\tilde{\phi}^{(L)}}$ and
$\Ket{\psi^{(L)}} = \alpha \Ket{\tilde{\psi}^{(L)}}$ where $\Ket{\tilde{\phi}^{(L)}}$
goes to $\Ket{\phi}$ when $L \rightarrow \infty$. If we knew $\alpha$, we would be
able to compute $\Braket{\tilde{\psi}^{(L)}|\tilde{\phi}^{(L)}} \rightarrow \beta$
as $L \rightarrow \infty$. Note that $\Ket{\phi}$ is a null state so that we cannot
simply normalize it, we thus need to find another way to get rid of this global
factor $\alpha$. This is achieved using a regularization of the Virasoro generators
on the lattice. Following Ref. \cite{KooSaleur}, for a general critical Hamiltonian
\begin{equation}
H = - \sum_{i=1}^{L-1} h_{i}
\end{equation}
with Fermi velocity $v_F$, we define a lattice version of the $L_n$'s
\begin{equation}
\displaystyle L^{(2N)}_n = \frac{L}{\pi} \left[ - \frac{1}{v_F} \sum_{i=1}^{L-1} (h_i - h_{\infty}) \cos \left( \frac{n i \pi}{L} \right) + \frac{1}{v_F^2} \sum_{i=1}^{L-2} \left[ h_i, h_{i+1} \right] \sin \left( \frac{n i \pi}{L} \right)\right] + \frac{c}{24} \delta_{n,0},
\label{eq_KooSaleur}
\end{equation}
where $h_{\infty}$ is the ground state expectation value of $h_i$. It is possible
to show that such expression provides a good way to define the Virasoro algebra
on the lattice in the case where $h_i=e_i$ is a Temperley-Lieb generator. In
particular, it is possible to measure the central charge or scalar products
of the continuum limit through the computation of Virasoro commutators. We
shall admit that this equation remains correct even if $h_i$ is a generator
of the dilute Temperley-Lieb algebra.
Note that these generators do not exactly satisfy the Virasoro algebra in
the continuum limit because of anomalies: the commutators of the scaling limit do not coincide in general
with the scaling limit of the commutators due to extra couplings to `non-scaling states' \cite{KooSaleur}.
While the problem can be solved using a double limit procedure, in practice,
the anomalies induce extremely small errors to the lattice measurements. For
practical purposes, formula~(\ref{eq_KooSaleur}) can thus be used naively,
even when commutators of multiple actions of Virasoro generators are involved.
Using formula (\ref{eq_KooSaleur}, we are thus able to construct a lattice version $A^{(L)}$
of the A operator that links $\phi$ and $\xi$. Moreover, the state $\Ket{\xi^{(L)}}$
is readily identified in the spectrum. We normalize it such that
$\Braket{\xi^{(L)}|\xi^{(L)}}=1$. If we assume that
$\Ket{\tilde{\phi}^{(L)}} = A^{(L)} \Ket{\xi^{(L)}}$ is a correct
lattice version of $\Ket{\phi}$, we are now ready to compute $\beta$.
Gathering all the pieces,
we define a lattice version of $\beta$ which does not depend on $\alpha$
\begin{equation}
\displaystyle \beta^{(L)} = \dfrac{\left| \Braket{\psi^{(L)}|A^{(L)} \xi^{(L)}} \right|^2 }{\Braket{\psi^{(L)}|\phi^{(L)}}}.
\label{eq_b_lattice}
\end{equation}
When $L \rightarrow \infty$, we check that
\begin{equation}
\displaystyle \lim_{L \rightarrow \infty} \beta^{(L)} = \lim_{L \rightarrow \infty} \dfrac{ |\alpha|^2 \left| \Braket{\tilde{\psi}^{(L)}|A^{(L)} \xi^{(L)}} \right|^2 }{|\alpha|^2 \Braket{\tilde{\psi}^{(L)}|\tilde{\phi}^{(L)}}} = \lim_{L \rightarrow \infty} \Braket{\tilde{\psi}^{(L)}|\tilde{\phi}^{(L)}}=\Braket{\psi|\phi}= \beta.
\end{equation}
We summarize our method to measure $\beta$ by the following steps:
\begin{enumerate}
\item Using exact diagonalization methods, find a Jordan basis for the first few excitations of $H$ on $L=2N$ sites.
\item Identify a Jordan cell in the spectrum of $H$ and normalize the states like in eq.~\eqref{eq_latticeH}.
\item Also identify the state $\Ket{\xi^{(L)}}$ and normalize it such that $\Braket{\xi^{(L)}|\xi^{(L)}}=1$ for the lattice scalar product.
\item Using Virasoro generators on the lattice~\eqref{eq_KooSaleur}, construct the operator $A^{(L)}$.
\item Compute $\beta^{(L)}$ using eq.~\eqref{eq_b_lattice}.
\end{enumerate}
The value of the indecomposability parameter $\beta = \lim_{L \rightarrow \infty} \beta^{(L)}$ is then computed using an extrapolation $\beta^{(L)} = \beta + A/L + B/L^2 + \dots$ We find numerically that $\beta^{(L)}$ does not depend on the chosen Temperley-Lieb representation (Loop, XXZ, or SUSY). However, it does depend on how
we realize the operator $A$ on the lattice\footnote{In general, there are infinitely many ways to realize the operator A on the lattice. In the case of $A=L_{-2}$
acting on the vacuum for example, one can use the trousers trick \cite{DJS}, or the Koo-Saleur formula~(\ref{eq_KooSaleur}) directly $A=L^{(2N)}_{-2}$, or even $A=L^{(2N)}_{-2}+ \alpha L^{(2N)}_{2}$ where $\alpha \in \mathbb{C}$. The values of $\beta^{(L)}$ computed using these different lattice realizations are different, although we expect them to yield the same result in the limit $N \rightarrow \infty$.}, this is why we were able to improve the results of Ref. \cite{DJS}.
\section{Numerical results}
We present in this section our numerical results for indecomposability parameters in different examples of LCFT.
\subsection{The case $x=1$ ($c=-2$): Dense Polymers}
We begin with the case of the XX(Z) chain of even length $2N$,
with $q=i$ so $c=-2$. The expansion of the partition function
in terms of dense loops describes dense polymers as the weight
for a closed loop is $n=0$. This theory is relevant for the
description of spanning trees \cite{spanningtrees} (up to a duality transformation).
It is also related with the description of abelian sandpile models,
although in the latter case different values of the indecomposability parameters have been found \cite{Sandpile2,Sandpile}.
In the SUSY language, this model also corresponds to $\mathfrak{gl}(m|m)$
(super)spin chains and to non-linear sigma models with target space
$\mathbb{CP}^{m-1|m}$ at $\theta = \pi$ \cite{RS1}. For $m=1$, we
get the $\mathfrak{gl}(1|1)$ spin chain which is a free fermion system.
Indeed, at $m=1$ the $U_q(\mathfrak{sl}_2)$ XX spin chain and the supersymmetric
$\mathfrak{gl}(1|1)$-invariant chain coincide and everything can be
reformulated in terms of free-fermion generators obeying
$\lbrace f_i,f_j \rbrace=0$ and $\lbrace f_i,f^\dagger_j \rbrace=(-1)^{i+1} \delta_{i,j}$.
Within this representation, the Temperley-Lieb generators read
\begin{equation}
\displaystyle e_i = (f^\dagger_i + f^\dagger_{i+1})(f_i + f_{i+1}).
\end{equation}
The corresponding continuum limit is a symplectic fermions
theory \cite{Kausch1, Kausch2} with action
\begin{equation}
\displaystyle S = \frac{1}{2 g_{\sigma}^2} \int \mathrm{d}^2 r \ \partial_{\mu} \xi^\dagger \partial^{\mu} \xi.
\end{equation}
This theory is probably one of the best understood LCFT. There
are $4$ different fields at level $0$ which are organized into
a diamond indecomposable module of $\mathfrak{gl}(1|1)$. All the Virasoro staggered modules
can be constructed as fermionic excitations of these states at
level $0$. A few indecomposability parameters have been computed
using the Kausch-Gaberdiel algorithm \cite{KauschGaberdiel}. One can
also readily compute the same parameters using the free fermion
representation.
We show here how to implement our lattice approach to this model. The Kac formula at $x=1$ reads
\begin{equation}
\displaystyle h_{r,s} = \frac{ \left(2r - s \right)^2 - 1}{8},
\end{equation}
and extending the generic results to the case $q=i$, we find
that the partition function of the $\mathfrak{gl}(1|1)$ spin chain is
\begin{equation}
Z = \sum_{j=0}^{\infty} (2j+1) \frac{q^{(2j-1)^2/8}-q^{(2j+3)^2/8} }{\eta(q)} = \sum_{j=1}^{\infty} j (2 \chi_{j,1}+\chi_{j+1,1}+\chi_{j-1,1}).
\end{equation}
Recall that $\chi_{j,1} = \chi_{1,1+2j} $ is the irreducible
character of the Virasoro simple module with $h=h_{j,1}=h_{1,1+2j}$.
Using the general results presented in section \ref{generalstructure} with $p=1$
and $j = \frac{1}{2} (\mathrm{mod} \ 1)$,
we see that there is no indecomposability for $L$ odd ($j$ half-integer)
so that the lattice Hamiltonian remains fully diagonalizable in this
case. For $j \in \mathbb{N}$, the staggered Virasoro modules have the
following subquotient structure and a basis in terms of fields
\begin{equation}
\mathcal{P}_{j}=
\begin{array}{ccccc}
&&\hskip-.7cm R_{j} &&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_{j-1} &&&\hskip-.3cm R_{j+1} \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_{j} &&
\end{array}
=
\begin{array}{ccccc}
&&\hskip-.7cm \psi^{(j)} &&\\
&\hskip-.2cm\swarrow&\searrow&\\
\xi^{(j)} &&&\hskip-.3cm \rho^{(j)} \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm \phi^{(j)}&&
\end{array}.
\end{equation}
The modules $\mathcal{P}_{j}$ are completely characterized by the
logarithmic couplings $\beta_{1,1+2j}$, and each module corresponds to a Jordan cell
\begin{align}
L_0 \phi^{(j)} & = h_{1,1+2j} \phi^{(j)} \\ \notag
L_0 \psi^{(j)} & = h_{1,1+2j} \psi^{(j)} + \phi^{(j)} \\ \notag
\phi^{(j)} & = A_j \xi^{(j)} \\ \notag
A_j^{\dagger} \psi^{(j)} & = \beta_{1,1+2j} \xi^{(j)} .
\end{align}
Note that we choose a different convention for the operator $A$ than
eq.~\eqref{eq_Aconv} in order to match \cite{KauschGaberdiel} as we
normalize it such that $A_j = L^{j-1}_{-1}+\dots$ Using this convention
and eq.~\eqref{b_formula}, we were able to conjecture a general formula
for the indecomposability parameter
\begin{equation}
\displaystyle \beta_{1,1 + 2j} = - \lim_{x \rightarrow 1} \frac{ \Braket{\xi |A^\dag A |\xi}}{h_{1,1+2j}-h_{1,1+2(j-1)}-(j-1)}= - \frac{[(2j-3)!]^2}{4^{j-2}}(j-1), \ \ j \in \mathbb{N}\setminus\{0,1\}.
\end{equation}
We checked this equation up to $j=16$ using eq.~\eqref{b_formula}.
Note also that in other cases can similar explicit formulae be obtained from
eq.~\eqref{b_formula}; we shall report on this point and compare
our results with the literature (see {\it e.g.} \cite{MathieuRidout1,Kytola}) in a separate publication \cite{AVJS}.
We can try to measure these numbers numerically using the method described
in section \ref{numericalmethod}. To do so, we need to construct the lattice
version of the $A_j$ operator using eq.~\eqref{eq_KooSaleur} with
$h_i=e_i$, $h_{\infty}= 2/ \pi$ and $v_F=2$. For $j=2,3,4$, these operators read
\begin{subequations}
\begin{eqnarray}
A_{2} &=& L_{-1} , \ \ \beta_{1,5}=-1 \\
A_{3} &=& L^2_{-1} - 2 L_{-2}, \ \ \beta_{1,7}=-18 \\
A_{4} &=& L^3_{-1} - 8 L_{-2} L_{-1} +12 L_{-3}, \ \ \beta_{1,9}=-2700
\end{eqnarray}
\end{subequations}
We measured these three parameters in the spin sectors $S_z=1, 2$ and $3$,
respectively. The results are presented in Tab. \ref{tab_b_dense_polymers},
in very good agreement with the theoretical expectation. We computed these
numbers using both supersymmetric (XX) and geometrical representations of
the Temperley-Lieb algebra and obtained the same values in finite size.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
$L = 2 N $ & $\beta_{1,5}$ & $\beta_{1,7}$ & $\beta_{1,9}$\\
\hline
8 & -0.937759 & -13.3574 &\\
10 & -0.959708 & -14.8908 & -1518.37 \\
12 & -0.971844 & -15.7936 & -1805.43 \\
14 & -0.979236 & -16.3612 & -2013.66 \\
16 & -0.984064 & -16.7384 & -2157.86 \\
18 & -0.987388 & -17.0006 & -2262.59 \\
20 & -0.989771 & -17.1898 & -2340.51 \\
22 & -0.991539 & -17.3304 & -2399.80 \\
24 & & & -2445.81 \\
\hline
$\infty$ & -1.0000 $\pm$ 0.0002 & -18.0(0) $\pm$ 0.05 & -27(00) $\pm$ 25\\
\hline
Exact & $-1$ & $-18$ & $-2700$\\
\hline
\end{tabular}
\end{center}
\caption{Measure of indecomposability parameters in the XXZ spin chain at $q=i$.}
\label{tab_b_dense_polymers}
\end{table}
\subsection{The cases $x=2$ ($c=0$) and $x=1/2$ ($c=-7$): Percolation and $\mathfrak{sl}(2|1)$ superspin chain}
We now deal with a slightly more complicated example, the antiferromagnetic
$\mathfrak{sl}(2|1)$ superspin chain. This chain is known to be equivalent
to the classical percolation problem, and arises naturally in the context
of the Spin Quantum Hall Effect \cite{SQHE}. The percolation problem has
$c=0$ and can also be formulated in terms of geometrical clusters or the XXZ
spin chain at $q=\mathrm{e}^{2i \pi/3}$. We also consider the ferromagnetic
version, with central charge $c=-7$.
\subsubsection{$\mathfrak{sl}(2|1)$-invariant spin chain}
We consider a chain of alternating fundamental and dual representations
of the Lie superalgebra $\mathfrak{sl}(2|1)$ with `Hilbert' space
$\mathcal{H} = (\square \otimes \bar{\square})^{\otimes N}$. For more
details about $\mathfrak{sl}(2|1)$ and its representation theory, we
refer the interested reader to the literature (see {\it e.g.} Ref. \cite{sl21}).
The Hilbert space on one specific site is spanned by three independent states
so the whole Hilbert space has dimension $3^{L}=9^{N}$. On every site, we introduce
two boson operators $[ b_{i,\sigma}, b^\dag_{j,\sigma'} ] = \delta_{ij} \delta_{\sigma\sigma'}$,
where $\sigma \in \{ \uparrow, \downarrow \}$, and one fermion
$\{ f_i, f_j^\dag \} = (-1)^{i+1} \delta_{ij}$. We add a constraint
to the system so that there cannot be more than one particle by site.
The representation on the site $i$ is $\mathbb C^3 \simeq \mathrm{Span} \{ f_i^\dag \Ket{0}, b_{i,\uparrow}^\dag \Ket{0}, b_{i,\downarrow}^\dag \Ket{0} \}$ and corresponds to the
fundamental $\square$ for $i$ odd and to the dual $\bar{\square}$
otherwise. The invariant coupling is chosen to be the Casimir in the
tensor product representation of the sites $i$ and $i+1$.
\begin{equation}
\displaystyle e_i = (b_{i+1,\downarrow}^\dag b_{i,\uparrow}^\dag + b_{i+1,\uparrow}^\dag b_{i,\downarrow}^\dag + (-1)^{i+1} f_{i+1}^\dag f_i^\dag) (b_{i,\uparrow} b_{i+1,\downarrow} + b_{i,\downarrow} b_{i+1,\uparrow} + (-1)^{i+1} f_i f_{i+1}).
\end{equation}
It furnishes a representation of the Temperley-Lieb algebra
with $n=1$. The Hamiltonian reads
\begin{equation}
\displaystyle H= \pm \sum_{i=1}^{L-1} e_i,
\end{equation}
where the minus sign corresponds to percolation ($c=0$) and
the plus sign to the ferromagnetic case ($c=-7$). There are
two good quantum numbers $S_z$ and $B$ that we can use to label the states
\begin{subequations}
\begin{eqnarray}
S_z &=& \frac{1}{2}\sum_{i=1}^{L} (b_{i,\uparrow}^\dag b_{i,\uparrow} - b_{i,\downarrow}^\dag b_{i,\downarrow}), \\
B &=& \sum_{i=1}^{L} \left( (-1)^{i+1} \frac{b_{i,\uparrow}^\dag b_{i,\uparrow} + b_{i,\downarrow}^\dag b_{i,\downarrow}}{2} + f^\dag_i f_i \right).
\end{eqnarray}
\end{subequations}
Finally, note that the low energy physics of this chain can
be described by a non-linear sigma model with target space $\mathbb{CP}^{1|1}$ at $\theta = \pi$ \cite{RS1}.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|}
\cline{1-6}
\multicolumn{4}{|c|}{Percolation ($q=\mathrm{e}^{i \pi/3}$)} & \multicolumn{2}{|c|}{Ferro $\mathfrak{sl}(2|1)$ ($q=\mathrm{e}^{2i \pi/3}$)} \\ \cline{1-6}
$L$ & $\beta_{1,4}$ & $ L$ & $\beta_{1,5}$ & $L$ & $\beta_{1,7}$\\
\hline
7 & -0.471874 & 8 & -0.609088 & 8 & -1.57616\\
9 & -0.476386 & 10 & -0.605858 & 10 & -1.77278\\
11 & -0.479983 & 12 & -0.606403 & 12 & -1.87138 \\
13 & -0.482724 & 14 & -0.607775 & 14 & -1.92567 \\
15 & -0.484837 & 16 & -0.609226 & 16 & -1.95770 \\
17 & -0.486503 & 18 & -0.610561 & 18 & -1.97762\\
19 & -0.487845 & 20 & -0.611738 & 20 & -1.99049\\
21 & -0.488946 & 22 & -0.612764 & 22 & -1.99906\\
\hline
$\infty$ & -0.5000 $\pm$ 0.0001 &$\infty$ & -0.6249 $\pm$ 0.0005 & $\infty$ & -2.00 $\pm$ 0.005 \\
\hline
\hline
Exact & -1/2 &Exact & -5/8 = 0.625 & Exact & $-2$\\
\hline
\end{tabular}
\end{center}
\caption{Measure of indecomposability parameters in $c=0$ and $c=-7$ theories.}
\label{tab_perco}
\end{table}
\subsubsection{Measure of indecomposability parameters}
First of all, let us focus on the first known indecomposability
parameter which concerns the stress energy tensor
in the percolation problem.
The Kac formula with $x=2$ reads
\begin{equation}
\displaystyle h_{r,s} = \frac{ \left(3r - 2s \right)^2 - 1}{24},
\end{equation}
and the values appearing in the spectrum are
\begin{equation}
\displaystyle h_{1,1+2j} = \frac{j(2j-1)}{3}.
\end{equation}
The partition function of the $q=\mathrm{e}^{i \pi/3}$ XXZ
spin chain with an even number of sites reads
\begin{equation}
\displaystyle Z = \sum_{j=0}^{\infty} (2j+1) \frac{q^{(4j-1)^2/24}-q^{(4j+5)^2/24} }{\eta(q)}.
\end{equation}
For the $\mathfrak{sl}(2|1)$-invariant chain, one
replaces $2j+1$ by $[2j+1]_{q'}$\footnote{We define the q-analog as $[n]_q = \frac{q^{n}-q^{-n}}{q-q^{-1}}$.} with $q'+q'^{-1}=3$ \cite{RS3,RS2}.
As we already argued in details in section \ref{subsec_catastrophe},
the stress energy tensor at $c=0$ must have a logarithmic partner $t(z)$
which corresponds to the field $\psi(z)$ with our notations.
It is now well admitted that the indecomposability parameter in this
case is $\beta_{1,5}=-5/8$. This number was also measured numerically
in Ref. \cite{DJS}. The Jordan cells equations in that case read
\begin{align}
L_0 \Ket{T} & = 2 \Ket{T} \\ \notag
L_0 \Ket{t} & = 2 \Ket{t} + \Ket{T} \\ \notag
\Ket{T} & = L_{-2} \Ket{0} \\ \notag
L_2 \Ket{t} & = b \Ket{0} .
\end{align}
Using the general structure (section \ref{generalstructure}), we see that
these states are organized into the following diamond structure
\begin{equation}
\mathcal{P}_2=
\begin{array}{ccccc}
&&\hskip-.7cm R_2&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_0&&&\hskip-.3cm R_3 \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_2&&
\end{array}
=
\begin{array}{ccccc}
&&\hskip-.7cm t&&\\
&\hskip-.2cm\swarrow&\searrow&\\
I&&&\hskip-.3cm \rho \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm T&&
\end{array}
\end{equation}
We measured $\beta_{1,5}$ for various Temperley-Lieb representations. In the
XXZ spin chain at $q=\mathrm{e}^{i\pi/3}$, the Jordan cell occurs
in the sector $S_z=0$, while in the $\mathfrak{sl}(2|1)$ SUSY case,
the Jordan cell for $T$ is to be found in the sector $(S_z,B)=(0,0)$.
We recall that the lattice indecomposability parameter given by eq.~\eqref{eq_b_lattice} does not depend on the chosen representation.
In the geometrical setup of percolation as dense loop gas with
fugacity $n=1$, the Hamiltonian remains fully diagonalizable and
there is no coefficient to measure here. Nevertheless, it is still
possible to slightly deform it \cite{DJS} so that Jordan cells appear,
and in this case we find the same values as in the other representations.
The results (Tab. \ref{tab_perco}) are in excellent agreement with
the prediction $b = -5/8$ and significantly improve the precision of the results
obtained from the trousers trick \cite{DJS}.
One can also measure the indecomposability parameters
$\beta_{1,1+2j}$ with $j$ half-integer from odd-length chains.
For instance, for $L$ odd, there is a Jordan cell at level 1
that corresponds to the Virasoro staggered module $\mathcal{P}_\frac{3}{2}$.
This Jordan cell occurs in the sector $S_z=1/2$, we call as usual
$\xi$ the unique state with $h=0$ in this sector, and $\psi$ and
$\phi$ the states with $h=1$. We normalize $\xi$ such
that $\Braket{\xi|\xi}=1$. The OPE formula~\eqref{b_formula}
gives a logarithmic coupling $\beta_{1,4} = \Braket{\phi | \psi} = -1/2$;
this value was also found by Mathieu and Ridout using different
methods \cite{MathieuRidout,MathieuRidout1}. We can measure this
coefficient on the lattice using the same method; once again,
the results are in excellent agreement with the theoretical
expectation (Tab. \ref{tab_perco}).
Another case of interest is $q=\mathrm{e}^{2 i \pi/3}$, with a
central charge $c=-7$. This model corresponds to an $\mathfrak{sl}(2|1)$
spin chain with ferromagnetic couplings. Graphical expansion of
the partition function yields a loop model with fugacity $n=-1$.
The conformal dimensions appearing in the spectrum are given
by the Kac formula for $x=1/2$
\begin{equation}
\displaystyle h_{1,1+2j} = \frac{j(j-2)}{3}.
\end{equation}
The first interesting Jordan cell arises at level 1, and
corresponds to the staggered module
\begin{equation}
\mathcal{P}_3=
\begin{array}{ccccc}
&&\hskip-.7cm R_3&&\\
&\hskip-.2cm\swarrow&\searrow&\\
R_2&&&\hskip-.3cm R_5 \\
&\hskip-.2cm\searrow&\swarrow&\\
&&\hskip-.7cm R_3&&
\end{array}.
\end{equation}
In this case, eq.~\eqref{b_formula} with $A=L_{-1}$ yields
$\beta_{1,7} = -2$. This coefficient was also computed by Kausch and Gaberdiel \cite{KauschGaberdiel}
thanks to the Nahm-Gaberdiel-Kausch algorithm.
This cell occurs in the sectors $S_z=-2,-1,1,2$
of the XXZ spin chain. Measures of $\beta_{1,7}$ in all these sectors
yield values in good agreement\footnote{Extrapolation are done
fitting data by $\beta+ A/L + B/L^2$, so the resulting fitting curves
need not be monotonic; this is particularly obvious in this case.}
with $\beta_{1,7} = -2$ (see Tab. \ref{tab_perco}). As in the
other cases, the lattice values of $\beta_{1,7}$ do not
depend on the chosen representation.
\begin{table}
\begin{center}
\begin{tabular}{|c|c|}
\hline
$L = 2 N $ & $\beta_{1,5}$ \\
\hline
8 & -1.26986 \\
10 & -1.29548 \\
12 & -1.31743 \\
14 & -1.33489 \\
16 & -1.34876 \\
18 & -1.35993 \\
20 & -1.36905 \\
22 & -1.37663 \\
\hline
$\infty$ & -1.4582(8) $\pm$ 0.0001 \\
\hline
Exact & $-35/24 \simeq -1.4583$ \\
\hline
\end{tabular}
\end{center}
\caption{Measure of $\beta_{1,5}$ in the XXZ spin chain at $q=\mathrm{e}^{i \pi/4}$.}
\label{tab_ising}
\end{table}
\subsection{The case $x=3$ ($c=\frac{1}{2}$): Logarithmic Ising model}
Let us consider the case $q=\mathrm{e}^{i \pi/4}$, which corresponds
to a central charge $c=\frac{1}{2}$. It corresponds to a dense loop
model with fugacity $n=\sqrt{2}$; this is of course the celebrated
(logarithmic) Ising model $\mathcal{LM}(3,4)$.
The spectrum is given by the Kac formula at $x=3$
\begin{equation}
\displaystyle h_{1,1+2j} = \frac{j(3j-1)}{4}.
\end{equation}
The partition function of the $U_{q=\mathrm{e}^{i \pi/4}}(\mathfrak{sl}_2)$ XXZ spin chain reads
\begin{equation}
\displaystyle Z = \sum_{j=0}^{\infty} (2j+1) \frac{q^{(1- 6 j)^2/48}-q^{(7+6j)^2/48} }{\eta (q)}.
\end{equation}
The algebraic structure of the continuum limit was studied in details
in section \ref{subsec_ising}. We measured numerically the
indecomposability parameter associated to the module~\eqref{eq_projisingex}.
The results are shown in Tab.~\ref{tab_ising} and are in excellent
agreement with the expected result $\beta_{1,5}=-35/24$ (see eq.~\eqref{eq_b_ising}).
\subsection{Another theory at $c=0$: $O(n \rightarrow 0)$ model and dilute polymers}
\begin{table}
\begin{center}
\begin{tabular}{|c|c|}
\hline
$L = 2 N $ & $\beta_{3,1}$ \\
\hline
4 & 0.021029 \\
6 & 0.145101 \\
8 & 0.276585 \\
10 & 0.382046 \\
12 & 0.463292 \\
14 & 0.526436 \\
\hline
$\infty$ & 0.9 $\pm$ 0.1 \\
\hline
Exact & 5/6 $\simeq$ 0.8333 $\dots$ \\
\hline
\end{tabular}
\end{center}
\caption{Measure of $\beta_{3,1}$ in the $O(n\rightarrow0)$ model. The convergence
is quite poor compared to the Temperley-Lieb case, but the result is not
so bad given that the precision we obtain is roughly the same as for the
first critical exponents. The result is consistent with the known value $b=\frac{5}{6}$.}
\label{tab_0n}
\end{table}
Finally, we present here an example of measure of indecomposability
parameters in `dilute' LCFTs. We study the $O(n \rightarrow 0)$ model,
which is known to be relevant for the physics of dilute polymers.
We can also formulate this model using supersymmetry, in terms of
$\mathfrak{osp}(2m|2m)$-invariant spin chain and to non-linear sigma
models with supersphere target space $S^{2m-1|2m}$ \cite{RS1}.
The partition function of the $S=1$ $U_q(\mathfrak{sl}_2)$-invariant chain
reads
\begin{equation}
\displaystyle Z = \sum_{j=0}^{\infty} (2j+1) \frac{q^{h_{1+2j,1}}-q^{h_{-1-2j,1}} }{P (q)} = \sum_{j=0}^{\infty} (2j+1) \frac{q^{(1+6j)^2/24}-q^{(5+6j)^2/24} }{\eta (q)},
\end{equation}
where the exponents appearing in the spectrum
are now in the first column $h_{1+2j,1}$ of the Kac table.
We are interested in the stress energy tensor $T$ in this theory,
which is primary in this case as the central charge is $c=0$. As
in the percolation problem, $T$ has a logarithmic partner that we
call $t$. However, while in the percolation theory $t$ had to be
somehow identified with the field $\Phi_{1,5}(z)$, it corresponds
here to $\Phi_{3,1}(z)$ with thus a completely different
indecomposability parameter $\beta_{3,1}=5/6$. We would like
to measure this coefficient from the lattice model defined
eq.~\eqref{hamOn}. Unfortunately, the finite-size convergence
of the exponents is very poor, even the central charge cannot
be properly measured in this case. This is probably due to the
absence of integrable K-matrices in our model; a similar phenomenon was observed
in Ref. \cite{KooSaleur2}. Of course, one could add K-matrices at the edges
in order to improve the convergence. Nevertheless, it is not clear to us how to adapt eq.~\eqref{eq_KooSaleur} in that case.
Note also that the Hilbert space is much
larger here so the accessible sizes are relatively small.
Nevertheless, we can still hope to deduce a rough estimate of
the indecomposability parameter $b$ for the stress energy
tensor here. Since the action of the operator $L_{2}$ on the
vacuum gives 0, we define the operator $A^{(L)}$ as
\begin{equation}
\displaystyle L^{(L)}_{2} + L^{(L)}_{-2} = - \frac{2L}{\pi v_F} \sum_{i=1}^{L-1} (h_i-h_{\infty}) \cos \left[ \frac{2 i \pi}{L} \right].
\end{equation}
Of course, we could also have used $A^{(L)}=L^{(L)}_{-2}$ but
as it turns out, the convergence is better with $L^{(L)}_{2} + L^{(L)}_{-2}$.
Measure of $b$ from this formula are shown in Fig.~\ref{tab_0n}.
Although the convergence is clearly not as good as for the
previous examples, the result is consistent with the value $5/6$.
\renewcommand{\arraystretch}{2.5}
\begin{table}
\begin{center}\small
\vspace{-2cm}
\begin{turn}{90}
\begin{tabular}{|c||c|c|c|c|c|c|c|}
\hline
\bf{Dense LCFTs}& $\beta_{1,3}$ & $\beta_{1,4}$ & $\beta_{1,5}$ & $\beta_{1,6}$ & $\beta_{1,7}$ & $\beta_{1,8}$ & $\beta_{1,9}$ \\
\hline
$\mathcal{LM}(1,5)$ ($x=\dfrac{1}{4}$) & $\proja$ & $\proja$ & $\projb$ & $\Diamond$ & $\Diamond$ & $\Diamond$ & $\Diamond$ \\
\hline
$\mathcal{LM}(1,4)$ ($x=\dfrac{1}{3}$) & $\proja$ & $\projb$ & $\Diamond$ & $\Diamond$ & $\Diamond$ & $\projb$ & $-3$ \\
\hline
$\mathcal{LM}(1,3)$ ($x=\dfrac{1}{2}$) & $\projb$ & $\Diamond$ & $\Diamond$ & $\projb$ & $-2$ & $8$ & $\projb$ \\
\hline
Dense Polymers $\mathcal{LM}(1,2)$ & $\Diamond$ & $\projb$ & $-1$ & $\projb$ & $-\dfrac{9}{2}$ & $\projb$ & $-\dfrac{75}{4}$ \\
\hline
Percolation $\mathcal{LM}(2,3)$ & $\projb$ & $-\dfrac{1}{2}$ & $-\dfrac{5}{8}$ & $\projb$ & $-\dfrac{35}{3}$ & $-\dfrac{13475}{216}$ & $\projb$ \\
\hline
Ising $\mathcal{LM}(3,4)$ & $\proja$ & $\projb$ & $-\dfrac{35}{24}$ & $-\dfrac{13475}{243}$ & $-\dfrac{49049}{17496}$ & $\projb$ & $-\dfrac{40415375}{944784}$ \\
\hline
Tricritical Ising $\mathcal{LM}(4,5)$ & $\proja$ & $\proja$ & $\projb$ & $-\dfrac{693}{100}$ & $-\dfrac{6114399291}{1078465600}$ & $-\dfrac{91820268514045071}{253871477862400}$ & $-\dfrac{21676129054392267}{1644513366760000}$ \\
\hline
3-state Potts $\mathcal{LM}(5,6)$ & $\proja$ & $\proja$ & $\proja$ & $\projb$ & $-\dfrac{676039}{59895}$ & $-\dfrac{22502936626745344}{562010429701125}$ & $-\dfrac{745930435583727415172151}{24227875374038666253125}$ \\
\hline
\hline
\bf{Dilute LCFTs} & $\beta_{3,1}$ & $\beta_{4,1}$ & $\beta_{5,1}$ & $\beta_{6,1}$ & $\beta_{7,1}$ & $\beta_{8,1}$ & $\beta_{9,1}$ \\
\hline
Dilute Polymers $\mathcal{LM}(2,3)$ & $\dfrac{5}{6}$ & $\projb$ & $\dfrac{67375}{676}$ & $\projb$ & $\dfrac{106462606250}{116550867}$ &$\projb$ & $\dfrac{141745038705442046875}{26751761399366832}$ \\
\hline
$O(n\rightarrow1)$ Ising $\mathcal{LM}(3,4)$ & $\projb$ & $\dfrac{175}{12}$ & $\dfrac{49049}{15552}$ & $\projb$ & $\dfrac{88913825}{229842}$ & $\dfrac{362318037060948052068359375}{6798093588426728083456}$ & $\projb$ \\
\hline
\end{tabular}
\end{turn}
\end{center}
\caption{
Indecomposability parameters of minimal chiral Logarithmic
Conformal Field Theories. We consider two different types of
LCFTs depending on whether the underlying loop models are in
a dilute or dense phase. Each parameter $\beta$ corresponds
to a diamond-shaped staggered Virasoro module. We use different
symbols when $\beta$ is not defined: $\proja$ standard indecomposable
module, $\projb$ irreducible (simple) module, or $\Diamond$ $L_0$-Jordan cell with no
interesting indecomposability parameter to compute.
}
\renewcommand{\arraystretch}{1.0}
\label{general_b_figure}
\end{table}
\subsection{Remarks on descendants}
We have seen that using the structure of staggered modules
over Virasoro which arise in a given theory, one can predict
the whole structure of Jordan cells in the continuum limit.
However, this does not take into account descendants: there
is a whole pyramid of Jordan cells associated with each
(primary) Jordan cell. To be more precise, if there is a
Jordan cell mixing two operators $\psi$ and $\phi$ with
parameter $\beta$, then we can expect Jordan cells for all
the descendants as is readily shown using the commutation
relations of the $L_n$'s. The resulting indecomposability
parameters are not independent and can be deduced from
the knowledge of $\beta$. For instance, let us consider
the operators $\psi^{(-n)} = L_{-n} \psi $ and
$\phi^{(-n)} = L_{-n} \phi$ in the case $A=L_{-2}$. In the basis
$( \psi^{(-n)},\phi^{(-n)} )$, $L_0$ reads
\renewcommand{\arraystretch}{1.0}
\begin{equation}
\displaystyle L_0 = \left( \begin{array}{cc} h+n & 1 \\ 0 & h+n \end{array} \right),
\end{equation}
where $h$ is the conformal weight of $\psi$ and $\phi$.
Let $\beta^{(-n)} = \Braket{\psi^{(-n)}|\phi^{(-n)}}$. Using
the Virasoro algebra, one can show that this coupling is given by
\begin{equation}
\displaystyle \beta^{(-n)} = \left( \frac{c}{12} n(n^2-1) + \frac{c}{2} \delta_{n,2} + 4 n \right) \beta.
\end{equation}
Of course, there are similar formulae for other kinds of descendants and for other $A$ operators.
We remark that these results are compatible with eq.~\eqref{b_formula}.
We measured indecomposability parameters for descendants in some cases
(results not shown here) and found a good agreement with the previous
considerations.
\section{Conclusion}
Pushing further the analysis of \cite{RS3}, we have shown in this paper that it is possible to investigate the
fine structure of indecomposable Virasoro modules in LCFTs using numerical analysis of
a certain type of lattice models.
Our method is general enough to be adapted to many cases,
and the precision reached is almost as good as for critical exponents.
We have restricted to the simplest type of boundary conditions for the LCFTs, but extension to more complicated cases is possible using more complicated lattice models, based for instance on the blob algebra \cite{blob1,blob2,blob3} (or the one and two-boundary Temperley-Lieb algebra). More interestingly maybe, we believe that the method can be extended to the periodic case as well and thus should provide a powerful tool
to investigate the structure of bulk LCFTs, where very little seems to be known at present. We will report on all these questions soon.
Going back to the values of the indecomposability parameters, we also argued that they can be
inferred from a simple heuristic argument relying on OPEs. This did not seem to be known, and suggests revisiting the bulk problem as well, by
systematically considering LCFTs as the limit of usual, non-logarithmic, CFTs. This will also be discussed elsewhere.
Finally, we summarize our results with a table of the first few indecomposability
parameters for the minimal logarithmic models $\mathcal{LM}(p,p')$.
We focus on the series $\mathcal{LM}(1,p)$ and $\mathcal{LM}(p,p+1)$,
and we consider the two versions `dense' and `dilute' of each theory.
Using the general structure of section \ref{generalstructure}
and eq.~\eqref{b_formula}, one can analyze the indecomposability
parameters in a systematic fashion \cite{AVJS}. The operators $A= L_{-n}+\dots$
are generated using the null-vector condition, and are normalized
as in eq.~\eqref{eq_Aconv}. The results are gathered in Tab.
\ref{general_b_figure}.
For a given dense ({\it resp.} dilute)
$\mathcal{LM}(p,p')$ theory, we denote $\beta_{1,1+2j}$
({\it resp.} $\beta_{1+2j,1}$) the logarithmic coupling associated
with the Jordan cell at level $h_{1,1+2j}$ ({\it resp.} $h_{1+2j,1}$).
Of course, it may happen that this Jordan cell does not exist, or
there may not be any interesting coefficient to measure (this is
the case for the first few Jordan cells in $\mathcal{LM}(1,p)$ theories), in which
cases we use different symbols. Some of the parameters we find are
quite complicated irreducible fractions, and the simplicity
of the results sometimes depends on the normalization choice for $A$.
Note that this table contains only a small fraction of the
couplings that eq.~\eqref{b_formula} allows to compute. In principle,
the formula~\eqref{b_formula} could be applied to obtain any indecomposability
parameter for a given theory. The limitation obviously comes from computing
Virasoro commutators. We shall report on all this in \cite{AVJS}.
\vspace{10pt}
{\bf Acknowledgments.} \hspace{5pt} We are grateful to R. Bondesan, A.M. Gainutdinov, A. Lazarescu, N. Read and V. Schomerus for fruitful
discussions. We also thank J. Dubail, M. Ridout and P. Ruelle for useful comments on the preprint version of this manuscript.
This work was supported by the Agence Nationale de la Recherche (grant ANR-10-BLAN-0414).
|
\section{Introduction}
In the standard gate-based model of quantum computation with $n$ qubits,
the initial state of the (closed) system is
represented by a $2^n$-dimensional complex vector $\psi$, and the
computation is described by a unitary time-evolution operator $U$.
Running the quantum computer implements the map
\begin{equation}
\psi \rightarrow U \psi.
\end{equation}
A key feature of this approach to quantum information processing is the
exponential classical information
storage capacity of the
wave function $\psi$. However, the number of one- and two-qubit gates
required to implement an arbitrary element
of the unitary group ${\rm U}(2^n)$ is at least $2^{2n}$, making the
construction of an arbitrary $U$ inefficient (using
elementary gates). The goal of quantum algorithm design is to compute
interesting cases of $U$ with a polynomial
number of elementary gates; two important examples are the algorithms
for factoring \cite{ShorSIAMJC97} and quantum
simulation \cite{LloydSci96}.
Although these algorithms are efficient, the number of gates required
for interesting applications is still large
and error-corrected qubits are therefore required \cite{Neilsen2000}.
Central to this gate (or quantum circuit) model of quantum computation
is the idea that a set of elementary one- and
two-qubit gates can be used to construct an arbitrary high-dimensional
$U$. But why should we use such a
decomposition in the first place? The reason is that the Hamiltonians
nature usually provides for us have one- and two-qubit terms
(one-body and two-body operators). Consider, for example, the following
somewhat generic solid-state
model of an array of $n$ coupled qubits,
\begin{eqnarray}
H_{\rm qc} = \sum_{i} \epsilon_i c_i^\dagger c_i
+\sum_{i < j}g_{ij} \, J_{\mu\nu} \, \sigma^\mu_i\otimes\sigma^\nu_j,
\label{tcsq model}
\end{eqnarray}
written in the basis of uncoupled-qubit eigenstates \cite{notationNote}.
Here $ i,j = 1, 2, \dots, n,$ and $\mu, \nu = x, y, z.$
The $\epsilon_i$ are the uncoupled qubit energies, $g_{ij}$ (with $i \neq j$)
are qubit-qubit interaction strengths, and $J_{\mu\nu}$ is a fixed dimensionless
tensor determined by the hardware.
It is clear from
(\ref{tcsq model}) that one- and two-qubit operations are naturally generated
in this system, and we know
(from universality \cite{LloydPRL95,BarencoPRA95}) that this is sufficient.
If the hardware also provided controllable terms of the form
\begin{eqnarray}
\sum_{i jk }g_{ijk} \, J_{\mu\nu\lambda}\sigma^\mu_i \otimes \sigma^\nu_j
\otimes \sigma^\lambda_k ,
\label{three-qubit interactionl}
\end{eqnarray}
then three-qubit gates could be used as primitives as well.
\section{QUANTUM COMPUTATION IN THE SES}
The idea we explore here is to perform a quantum computation in the
$n$-dimensional single-excitation
subspace (SES) of the full $2^n$-dimensional Hilbert space. This is the
subspace spanned by the computational
basis states
\begin{equation}
|m) \equiv c_m^\dagger |00 \cdots 0\rangle = |0 \cdots 1_m \cdots 0\rangle,
\end{equation}
with $m=1,2,\dots,n.$ Experimentally, it is possible to prepare the quantum
computer in the SES, and it will
remain there with high probability as long as (i) the coupling strengths
$|g_{ij}|$ are much smaller than the
$\epsilon_i$; and (ii) certain single-qubit operations such as $\pi$ pulses
are not used (however, $2 \pi$ pulses are permitted and turn out to be
extremely useful, and $\pi$ pulses can be used to prepare SES states from the
ground state $|00 \cdots 0\rangle$).
The advantage of working in the SES can be understood from the following
expression \cite{PritchettPre10,matrixelementsNote} for the
SES matrix elements of model (\ref{tcsq model}),
\begin{widetext}
\begin{equation}
\big( m \big| H_{\rm qc} \big|m' \big) = \bigg[ \epsilon_m
- 2J^{zz} \big( \sum_{k<m} g_{km} + \sum_{k>m} g_{mk} \big)
+ J^{zz} \big(\sum_{i<j} g_{ij}\big) \bigg] \delta_{mm'}
+ \bigg[J^{xx} + J^{yy} - i (J^{xy}-J^{yx}) \bigg] g_{mm'}.
\label{ses matrix elements}
\end{equation}
\end{widetext}
Therefore, we have a high degree of control over the part of the Hamiltonian
in the SES.
For example, in the simple case of an array of qubits coupled with a tunable
$\sigma^{x}_i \otimes \sigma^{x}_j$ exchange interaction,
\begin{equation}
\big( m \big| H_{\rm qc} \big|m' \big) = \epsilon_m \delta_{mm'} + g_{mm'},
\label{xx case matrix elements}
\end{equation}
so the diagonal and off-diagonal elements are directly and independently
controlled by the qubit energies and coupling strengths, respectively.
Because of this high degree of controllability, $n$-dimensional unitary
operations can be carried out in a single step, bypassing the need to
decompose into elementary gates.
This property also enables the direct quantum simulation of real but otherwise
arbitrary time-dependent Hamiltonians, allowing a polynomial quantum speedup
with sub-threshold-fidelity qubits \cite{PritchettPre10}.
A quantum computer operating in the SES mode described here requires every
qubit to be tunably coupled to
the others, as implied by model (\ref{tcsq model}). This requires $n(n-1)/2$
coupling wires and associated circuits.
Consider, for example, the recently demonstrated tunable inductive coupler
for superconducting phase
qubits \cite{PintoPRB10,BialczakPRL11} (other tunable couplers for superconducting
qubits have also been
successfully demonstrated \cite{vanderPloegPRL07,NiskanenSci07,YamamotoPRB08,AllmanPRL10}).
The circuit diagrams for a single phase qubit ``q" and single coupler
``c" are illustrated in Fig.~\ref{coupler figure},
where the crossed boxes represent Josephson junctions. In terms of these
elements, a possible layout for a fully connected array is shown in
Fig.~\ref{layout figure} for the case
of $n=5$.
The SES computation method is not scalable, because it requires a qubit
for every Hilbert space
dimension used. However the advantage is that high-dimensional unitary
operations can be carried
out in a single step without the need to decompose the operation into
elementary gates. This
property allows processors of even modest sizes to perform quantum computations
that would otherwise (using a gate-based approach) require thousands of elementary
gates and therefore error-corrected qubits. In the remainder of this paper,
we illustrate the SES computation
method by applying it to Grover's search algorithm \cite{GroverPRL97}.
\section{application to Grover's search algorithm}
Here we consider the search for a single marked item in a database of size $n$,
where $n$ is the
number of qubits in the SES processor (\ref{tcsq model}). Within the SES,
Grover's algorithm \cite{GroverPRL97} is represented by the following control sequence,
\BEq
\big({W}{\rm O}_{m'}\big)^{n_{\rm Grover}}U_{\rm unif}
|1) \approx |m'),
\EEq
where $U_{\rm unif}$ generates the uniform superposition of all SES states $|m)$,
$m=1,2,\dots,n$. Here ${\rm O}_{m'}$ is the oracle corresponding to the marked
state $|m')$, $W$ is Grover's inversion operator, and $n_{\rm Grover} \approx (\pi/4)\sqrt{n}$
is the number of iteration steps of the algorithm. In the following sections we
show how each of these unitary operators can be generated in a {\it single} step.
\subsection{Preparation of the uniform state}
The uniform superposition state,
\BEq
|\psi_{\rm unif}) \equiv \frac{1}{\sqrt{n}}\sum_{m=1}^n |m) =
\frac{1}{\sqrt{n}}
\begin{bmatrix}
1&1&1&1&\dots
\end{bmatrix}^{\rm T},
\EEq
can be generated from the state
$|1) =
\begin{bmatrix}
1 & 0 & 0 & 0 & \dots
\end{bmatrix}^{\rm T}$,
where $^{\rm T}$ stands for transposition, through
\BEqA
\label{eq:psi_unif}
|\psi_{\rm unif})=ie^{i\a_{\rm unif}} e^{-iH_{\rm unif}t_{\rm unif}}|1),
\label{uniform state identity}
\EEqA
with
\begin{equation}
\a_{\rm unif} = \pi/(2\sqrt{n})
\end{equation}
and
\begin{equation}
t_{\rm unif} = \pi/(2g\sqrt{n}),
\end{equation}
using the SES Hamiltonian
\BEqA
H_{\rm unif} &=&
g \begin{bmatrix}
2& 1 & 1 & 1&\dots \cr
1& 0 & 0 & 0&\dots \cr
1& 0 & 0 & 0&\dots \cr
1& 0 & 0 & 0&\dots \cr
\vdots&\vdots &\vdots &\vdots &\ddots \cr
\end{bmatrix}.
\EEqA
This can be seen by noticing that the spectrum of $H_{\rm unif}$ and the
transformation $S$ (whose columns are shown unnormalized here for notational
simplicity) that diagonalizes $H_{\rm unif}$ via
$H_{\rm unif}^{\rm diag}=S^{\dagger} H_{\rm unif}S$ are given by
\BEq
E = 1\mp\sqrt{N}, 0, \dots, 0,
\EEq
and
\BEq
S = \begin{pmatrix}
1-\sqrt{N} &1+\sqrt{N}& 0&0&0&\dots &0\cr
1& 1 &-1 &-1&-1&\dots &-1\cr
1 &1& 1 &0&0&\dots &0\cr
1&1&0&1&0&\dots &0 \cr
1&1&0&0&1&\dots &0\cr
\vdots&\vdots &\vdots &\vdots &\vdots &\ddots &\vdots \cr
1&1&0&0&0&\dots&1
\end{pmatrix},
\EEq
respectively. Direct exponentiation then immediately leads to Eq. (\ref{eq:psi_unif}).
The initial state $|1)$ required in (\ref{uniform state identity}) is easily
prepared from the
ground state via a $\pi$ pulse.
\begin{figure}
\includegraphics[width=6.0cm]{fig1.eps}
\caption{
Definitions of two-terminal phase qubit circuit element ``q" and coupler circuit
element``c", expressed
in terms of their conventional circuit diagrams involving inductors, capacitors,
current biases, and
Josephson junctions (crossed boxes) \cite{PintoPRB10,BialczakPRL11}.}
\label{coupler figure}
\end{figure}
\begin{figure}
\includegraphics[width=7.0cm]{fig2.eps}
\caption{
(color online) Design for a fully connected network of $n$ Josephson
phase qubits. Each
qubit ``q" is suplied with a wire to which $n-1$ coupler leads are attached.
The case of $n=5$
is shown. The proposed design minimizes crossovers.}
\label{layout figure}
\end{figure}
\subsection{Single-step oracle and $W$ operators}
The oracle (corresponding to the marked state with $m'=3$, for instance), is
\BEqA
{\rm O} &=& 1 - 2|m')(m'| =
\begin{bmatrix}
1& 0 & 0 & 0&\dots \cr
0& 1 & 0 & 0&\dots \cr
0& 0 &-1 & 0&\dots \cr
0& 0 & 0 &1 &\dots \cr
\vdots&\vdots &\vdots &\vdots &\ddots \cr
\end{bmatrix},
\EEqA
which can be generated via
\BEq
W = e^{-i H_{\rm O} t_{\rm O}}, \quad t_{\rm O} = \pi/ \delta \epsilon,
\EEq
using the SES Hamiltonian
\BEqA
H_{\rm O} &=&
\begin{bmatrix}
0& 0 & 0 & 0&\dots \cr
0& 0 & 0 & 0&\dots \cr
0& 0 &-\delta \epsilon & 0&\dots \cr
0& 0 & 0 & 0 &\dots \cr
\vdots&\vdots &\vdots &\vdots &\ddots \cr
\end{bmatrix}.
\EEqA
$\delta \epsilon$ is the detuning of qubit $m'$ from the remaining qubits,
all having common frequency $\epsilon$.
This operation is simply a $2\pi$ rotation on the qubit associated with the
marked state $m'$. (Although we have
implemented it as a $z$ rotation, an $x$ or $y$ rotation would work equally
well.)
Notice that in the limit
$g \ll \epsilon$, which guarantees good isolation of the SES, the oracle can
be made arbitrarily fast if we choose sufficiently large
detuning $\delta \epsilon$.
Finally, the $W$ operator,
\BEqA
{\rm W} &=& 2|\psi_{\rm unif})(\psi_{\rm unif}| - 1
\nonumber \\
&=&
\frac{1}{n}\begin{bmatrix}
2-n& 2 & 2 & 2&\dots \cr
2& 2-n & 2 & 2&\dots \cr
2& 2 &2-n & 2&\dots \cr
2& 2 & 2 &2-n &\dots \cr
\vdots&\vdots &\vdots &\vdots &\ddots \cr
\end{bmatrix},
\EEqA
can be generated via
\BEqA
{\rm W} = e^{-i \a_{\rm W}}e^{-i H_{\rm W} t_{\rm W}},
\nonumber \\
\a_{\rm W} = (1-n)\pi/n, \quad
t_{\rm W} = \pi/(ng),
\EEqA
with
\BEqA
H_{\rm W} &=&
g \begin{bmatrix}
0& 1 & 1 & 1&\dots \cr
1& 0 & 1 & 1&\dots \cr
1& 1 & 0 & 1&\dots \cr
1& 1 & 1 & 0 &\dots \cr
\vdots&\vdots &\vdots &\vdots &\ddots \cr
\end{bmatrix}.
\EEqA
\subsection{Comparison with gate-based computation}
Here we compare the SES and conventional gate-based approaches for
a $n=256$ item search.
The search requires 12 iterations.
In the SES case, we choose a weak, $g/2\pi = 1.25$ MHz coupling, which
guarantees that the $W$ operation is at least 1.5 ns long.
We also choose a strong, $\delta \epsilon/2\pi = 100$ MHz
qubit detuning to guarantee that the oracle is sufficiently fast.
Then $t_{\rm unif} = 12.5$ ns,
$t_{\rm O} = 5$ ns,
$t_{\rm W} = 1.56$ ns, and the total duration of the algorithm
is about 100 ns, which is experimentally practical with current
superconducting architectures.
The SES estimate should be contrasted with the conventional approach,
which requires an input register of 8 qubits, an output register of 1
qubit, plus 7 additional ancilla qubits (to make the multiply controlled
gates more efficient).
The corresponding search oracle involves an 8-fold CNOT gate,
${\rm C}^8{\rm NOT}$, each of which can be made out of 85 two-qubit
CNOTs (using 7 ancillas) \cite{Neilsen2000}. The $W$ operator involves
a 7-fold controlled-Z gate, ${\rm C}^7{\rm Z}$, which uses 73 standard
CNOTs (plus 6 ancillas). Thus, a conventional 256-item gate-based Grover
search requires 158 CNOT gates per search step.
The full algorithm then contains nearly 2000 CNOT gates, plus local
rotations.
\section{discussion}
We have described an approach to superconducting quantum computation in which
the computation is carried out in the single-excitation subspace of the
full Hilbert space. Relative
to the standard gate-based approach, the SES method requires exponentially
more qubits and is therefore
nonscalable. The hardware requirements are also highly demanding: The fully
connected array of $n$
qubits requires $n(n-1)/2$ coupling wires and tuning circuits.
But the SES approach is much
more time efficient, permitting far larger computations than currently
possible using today's
sub-theshold-fidelity qubits in the standard way. This was illustrated
above for Grover's search
algorithm. We would expect a similar result for Shor's algorithm, but
note that practical factoring applications
would require impossibly large processor sizes. General purpose
time-dependent quantum simulation
can also be carried out in the SES, allowing an $n^3$ polynomial
quantum speedup \cite{PritchettPre10}.
However, when a large scale error-corrected quantum computer
eventually becomes available,
the gate-based approach will perform better than the SES method.
\newpage
\begin{acknowledgments}
This work was supported by the National Science Foundation under CDI grant DMR-1029764.
It is a pleasure to thank
Joydip Ghosh, John Martinis, Emily Pritchett, Andrew Sornborger and Phillip Stancil
for useful discussions.
\end{acknowledgments}
|
\section{Introduction}
In the previous work \cite{Cheng04}, various $P \rightarrow M$ form factors,
where $P$ represents a heavy pseudoscalar meson $(D$ or $B)$, and $M$
represents either $s$-wave or low-lying $p$-wave meson, were calculated
within the framework of the covariant light-front (CLF)
approach. This formalism preserves the Lorentz covariance in the
light-front framework and has been applied successfully to describe
various properties of pseudoscalar and vector mesons \cite{Jaus90,Jaus91,Jaus99}. The
analysis of the covariant light-front quark model to transitions of the
charmed and bottom mesons was extended to even parity, $p$-wave mesons \cite{Cheng04}.
Recently, the CLF approach has also been used to the studies of the
quarkonia \cite{Shen08, Hwang07}, the $p$-wave meson emitting decays of the bottom mesons
\cite{Cheng10} and the $B_c$ system \cite{Wang09} and so on.
In the present work, we update our results for $D$ and $B$ meson form factors, and
extend this analysis to determine the form factors appearing in the
$D_s , B_s\rightarrow M$ transitions, and to the flavor-diagonal
final state mesons $M$. Experimental measurements of the decays of the $\tau$ lepton,
pseudoscalar and vector mesons are employed to determine the decay
constants, which in turn fix the shape parameters, $\beta $, of the respective mesons.
For a few cases, the decay constants estimated by lattice calculations
have been used for this purpose. We have now used the improved estimation
of the $K_{1A}$ and $K_{1B}$ mixing angle, where $K_{1A}$ and $K_{1B}$
are the $^3 P_1 $ and $^1 P_1 $ states of $K_1$, respectively, which are
related to the physical $K_1 (1270)$ and $K_1 (1400)$ states.
We then study transitions of the heavy flavor pseudoscalar mesons to pseudoscalar
mesons ($P$), vector mesons ($V$), scalar mesons ($S$) and axial vector mesons ($A$)
within the CLF model. Numerical results of the form factors for these
transitions and their momentum dependence are presented in
detail. In particular, all the form factors for heavy-to-light and
heavy-to-heavy transitions for charmed mesons $(D, D_s)$ and bottom mesons $(B, B_s)$
are calculated. Further, their sensitivity to uncertainties of $\beta$
parameters of the initial as well as of the
final mesons is investigated separately. Theoretically, the Isgur-Scora-Grinstein-Wise
(ISGW) quark model \cite{ISGW, IW89} has been the only model for a long time
that could provide a systematical estimate of the transition
of a ground-state $s$-wave meson to a low-lying $p$-wave
meson. However, this model is based on the nonrelativistic constituent quark
picture. We have earlier pointed out
\cite{Cheng04} that relativistic effects could manifest
in heavy-to-light transitions at maximum recoil where the final-state meson can be
highly relativistic. For example, the
$B\rightarrow a_1$ form factor $V^{Ba_1}_0 (0)$ is found to be 0.13
in the relativistic light-front model \cite{Cheng04}, while it is as big as
1.01 in the ISGW model \cite{ISGW}.
Hence there is no reason to expect that the
nonrelativistic quark model is still applicable there, though in the
improved version of the model (ISGW2) \cite{ISGW2} a number of improvements,
such as the constraints imposed by heavy quark symmetry and
hyperfine distortions of wave functions have been incorporated.
We believe that the CLF quark model can provide
useful and reliable information on $B \rightarrow M$ transitions
particularly at maximum recoil.
The paper is organized as follows. The basic features of the covariant
light-front (CLF) model are recapitulated in Sec. II. In Sec. III, decay constants
are presented in the CLF model. Available experimental
measurements for various decays are used to determine decay constants,
which in turn are used to fix $\beta$ parameters of the CLF model.
Sometimes, lattice predictions for few decay constants are also used for this purpose.
In Sec. IV, the analysis of form factors appearing for transitions from pseudoscalar mesons
to $s$-wave mesons (pseudoscalar or vector) and $p$-wave mesons
(scalar and axial vector) is given. In Sec. V, numerical results are presented
for these form factors and their $q^2$ - dependence taking proper
inclusions of uncertainties in the shape parameter, $\beta$. Summary and conclusions
are given in Sec. VI.
\section{Formalism of a covariant light-front model}
\begin{figure}[t!]
\centerline{
{\epsfxsize2.1 in \epsffile{A.eps}}
\hspace{1cm}
{\epsfxsize3 in \epsffile{B.eps}}}
\centerline{\,\,\,\,\,(a)\hspace{6.2cm}(b)} \vskip0.2cm
\caption{Feynman diagrams for (a) meson decay and (b) meson
transition amplitudes, where $P^{\prime({\prime\prime})}$ is the incoming
(outgoing) meson momentum, $p^{\prime({\prime\prime})}_1$ is the quark
momentum, $p_2$ is the anti-quark momentum and $X$ denotes the
corresponding $V-A$ current vertex.}\label{fig:feyn
\end{figure}
In the conventional light-front framework, the constituent quarks
of the meson are required to be on their mass shells
and various physical quantities are
extracted from the plus component of the corresponding current
matrix elements. However, this procedure will miss the zero-mode
effects and render the matrix elements non-covariant. Jaus
\cite{Jaus90,Jaus91} has proposed a covariant light-front approach that
permits a systematical way of dealing with the zero mode
contributions. Physical quantities such as the decay constants and
form factors can be calculated in terms of Feynman momentum loop
integrals which are manifestly covariant. This of course means
that the constituent quarks of the bound state are off-shell. In
principle, this covariant approach will be useful if the vertex
functions can be determined by solving the QCD bound state
equation. In practice, we would have to be contended with the
phenomenological vertex functions such as those employed in the
conventional light-front model. Therefore, using the light-front
decomposition of the Feynman loop momentum, say $p_\mu$, and
integrating out the minus component of the loop momentum $p^-$,
one goes from the covariant calculation to the light-front one.
Moreover, the antiquark is forced to be on its mass shell after
$p^-$ integration. Consequently, one can replace the covariant
vertex functions by the phenomenological light-front ones.
To begin with, we consider decay and transition amplitudes given
by one-loop diagrams as shown in Fig.~\ref{fig:feyn} for the decay
constants and form factors of ground-state $s$-wave mesons and
low-lying $p$-wave mesons. We follow the approach of
\cite{Jaus99, Cheng04} and use the same notation. The incoming (outgoing)
meson has the momentum $P^{\prime({\prime\prime})}=p_1^{\prime({\prime\prime})}+p_2$,
where $p_1^{\prime({\prime\prime})}$ and $p_2$ are the momenta of the
off-shell quark and antiquark, respectively, with masses
$m_1^{\prime({\prime\prime})}$ and $m_2$. These momenta can be expressed in
terms of the internal variables $(x_i, p_\bot^\prime)$,
\begin{eqnarray}
p_{1,2}^{\prime+}=x_{1,2} P^{\prime +},\qquad
p^\prime_{1,2\bot}=x_{1,2} P^\prime_\bot\pm p^\prime_\bot,
\end{eqnarray}
with $x_1+x_2=1$. Note that we use $P^{\prime}=(P^{\prime +},
P^{\prime -}, P^\prime_\bot)$, where $P^{\prime\pm}=P^{\prime0}\pm
P^{\prime3}$, so that $P^{\prime 2}=P^{\prime +}P^{\prime
-}-P^{\prime 2}_\bot$.
In the covariant light-front approach, total four momentum is
conserved at each vertex where quarks and antiquarks are
off-shell. These differ from the conventional light-front approach
(see, for example ~\cite{Jaus91, Cheng97}) where the plus and
transverse components of momentum are conserved, and quarks as
well as antiquarks are on-shell.
It is useful to define some internal quantities for on-shell quarks:
\begin{eqnarray} \label{eq:internalQ}
M^{\prime2}_0
&=&(e^\prime_1+e_2)^2=\frac{p^{\prime2}_\bot+m_1^{\prime2}}
{x_1}+\frac{p^{\prime2}_{\bot}+m_2^2}{x_2},\quad\quad
\widetilde M^\prime_0=\sqrt{M_0^{\prime2}-(m^\prime_1-m_2)^2},
\nonumber\\
e^{(\prime)}_i
&=&\sqrt{m^{(\prime)2}_i+p^{\prime2}_\bot+p^{\prime2}_z},\quad\qquad
p^\prime_z=\frac{x_2 M^\prime_0}{2}-\frac{m_2^2+p^{\prime2}_\bot}{2 x_2 M^\prime_0}.
\end{eqnarray}
Here $M^{\prime2}_0$ can be interpreted as the kinetic invariant
mass squared of the incoming $q\bar q$ system, and $e_i$ the
energy of the quark $i$.
\begin{table}[b]
\caption{\label{tab:feyn} Feynman rules for the vertices
($i\Gamma^\prime_M$) of the incoming mesons-quark-antiquark, where
$p^\prime_1$ and $p_2$ are the quark and antiquark momenta,
respectively. Under the contour integrals to be discussed below,
$H^\prime_M$ and $W^\prime_M$ are reduced to $h^\prime_M$ and
$w^\prime_M$, respectively, whose expressions are given by
Eq.~(\ref{eq:h}). Note that for outgoing mesons, we shall use
$i(\gamma_0\Gamma^{\prime\dagger}_M\gamma_0)$ for the
corresponding vertices.}
\begin{tabular}{|c| c|}
\hline
$M\,(^{2S+1}L_J) $
&$i\Gamma^\prime_M$
\\
\hline
pseudoscalar ($^1S_0$)
&$H^\prime_P\gamma_5$
\\
vector ($^3S_1$)
&$i H^\prime_V [\gamma_\mu-\frac{1}{W^\prime_V}(p^\prime_1-p_2)_\mu]$
\\
scalar ($^3P_0$)
&$-i H^\prime_S$
\\
axial ($^3 P_1$)
&$-i H^\prime_{^3\!A}[\gamma_\mu+\frac{1}{W^\prime_{^3\!A}}(p^\prime_1-p_2)_\mu]\gamma_5$
\\
axial ($^1 P_1$)
&$-i H^\prime_{^1\!A} [\frac{1}{W^\prime_{^1\!A}}(p^\prime_1-p_2)_\mu]\gamma_5$\\
\hline
\end{tabular}
\end{table}
It has been shown in \cite{CM69} that one can pass to the
light-front approach by integrating out the $p^-$ component of the
internal momentum in covariant Feynman momentum loop integrals.
We need Feynman rules for the meson-quark-antiquark vertices to
calculate the amplitudes shown in Fig.~1. These Feynman rules for
vertices ($i\Gamma^\prime_M$) of ground-state $s$-wave mesons and
low-lying $p$-wave mesons are summarized in Table~\ref{tab:feyn}.
Next, we shall find the decay constants in the covariant light-front approach.
\section{Decay constants}
The decay constants for $J=0,1$ mesons are defined by the matrix
elements
\begin{eqnarray} \label{eq:AM}
\langle 0|A_\mu|P(P^\prime)\rangle &\equiv& {\cal A}^{P}_\mu=i f_P P^\prime_\mu ,\qquad
\langle 0|V_\mu|S(P^\prime)\rangle\equiv {\cal A}^{S}_\mu= f_S P^\prime_\mu ,
\\
\langle 0|V_\mu|V(P^\prime,\varepsilon')\rangle &\equiv& {\cal A}^{V}_\mu=M^\prime_V
f_V\varepsilon^\prime_\mu,\quad
\langle 0|A_\mu|\,^{3(1)}\!A(P^\prime,\varepsilon')\rangle \equiv {\cal A}^{^3\!A(^1\!A)}_\mu
=M^\prime_{^3\!A(^1\!A)} f_{^3\!A(^1\!A)}\varepsilon^\prime_\mu, \nonumber
\end{eqnarray}
where the $^{2S+1} L_J= {}^1S_0$, $^3P_0$, $^3S_1$, $^3P_1$ and
$^1P_1$ states of $q_1^\prime \bar q_2$ mesons are
denoted by $P$, $S$, $V$, $^3\!A$ and $^1\!A$, respectively.
It is
useful to note that in the SU(N)-flavor limit ($m_1^\prime=m_2$)
we should have vanishing $f_S$ and $f_{^1\!A}$. The former can be
seen by applying equations of motion to the matrix element of the
scalar resonance in Eq. (\ref{eq:AM}) to obtain
\begin{eqnarray} \label{eq:Seom}
m_S^2f_S=\,i(m'_1-m_2)\langle 0|\bar q_1q_2|S\rangle.
\end{eqnarray}
The latter is based on the argument that the light $^3P_1$ and
$^1P_1$ states transfer under charge conjugation as
\begin{eqnarray}
M_a^b(^3P_1) \to M_b^a(^3P_1), \qquad M_a^b(^1P_1) \to
-M_b^a(^1P_1),~~~(a=1,2,3),
\end{eqnarray}
where the light axial-vector mesons are represented by a $3\times
3$ matrix. Since the weak axial-vector current transfers as
$(A_\mu)_a^b\to (A_\mu)_b^a$ under charge conjugation, it is clear
that the decay constant of the $^1P_1$ meson vanishes in the SU(3)
limit \cite{Suzuki}. This argument can be generalized to heavy
axial-vector mesons. In fact, under similar charge conjugation
argument [$(V_\mu)_a^b\to -(V_\mu)_b^a$, $M_a^b(^3P_0) \to
M_b^a(^3P_0)$] one can also prove the vanishing of $f_S$ in the
SU(N) limit.
Furthermore, in the heavy quark limit ($m_1^\prime\to\infty$), the
heavy quark spin $s_Q$ decouples from the other degrees of freedom
so that $s_Q$ and the total angular momentum of the light
antiquark $j$ are separately good quantum numbers. Hence, it is
more convenient to use the $L^j_J=P^{3/2}_2$, $P^{3/2}_1$,
$P^{1/2}_1$ and $P^{1/2}_0$ basis. It is obvious that the first
and the last of these states are $^3P_2$ and $^3P_0$,
respectively, while \cite{IW91}
\begin{equation} \label{eq:Phalf}
\left|P^{3/2}_1\right\rangle=\sqrt{\frac{2}{3}}\,\left|^1P_1\right\rangle
+{1\over \sqrt{3}}\,\left|^3P_1\right\rangle,\qquad
\left|P^{1/2}_1\right\rangle={1\over \sqrt{3}}\,\left|^1P_1\right\rangle
-\sqrt{\frac{2}{3}}\,\left|^3P_1\right\rangle.
\end{equation}
Heavy quark symmetry (HQS) requires~\cite{IW89,HQfrules}
\begin{equation} \label{eq:HQSf}
f_V=f_P,\qquad
f_{A^{1/2}}=f_S,\qquad
f_{A^{3/2}}=0,
\end{equation}
where we have denoted the $P^{1/2}_1$ and $P^{3/2}_1$ states by
$A^{1/2}$ and $A^{3/2}$, respectively.
These relations in the above equation can be understood from the
fact that $(S^{1/2}_0,S^{1/2}_1)$, $(P_0^{1/2},P_1^{1/2})$ and
$(P_1^{3/2},P_2^{3/2})$ form three doublets in the HQ limit and
that the tensor meson cannot be induced from the $V-A$ current.
Following the procedure described in \cite{Jaus99, Cheng04},
we now evaluate meson decay constants through the following formulas:
\begin{eqnarray}
f_P=\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot \frac{h^\prime_P}{x_1
x_2 (M^{\prime2}-M^{\prime2}_0)}4(m_1^\prime x_2+m_2 x_1),
\label{eq:fP}
\end{eqnarray}
\begin{eqnarray}
f_V&=&\frac{N_c}{4\pi^3M^\prime}\int dx_2 d^2p^\prime_\bot \frac{h^\prime_V}{x_1
x_2 (M^{\prime2}-M^{\prime2}_0)}
\nonumber\\
&&\qquad\qquad\times\left[x_1 M^{\prime2}_0-m_1^\prime (m_1^\prime-m_2)-p^{\prime2}_\bot
+\frac{m_1^\prime+m_2}{w^\prime_V}\,p^{\prime2}_\bot \right],
\label{eq:fV}
\end{eqnarray}
\begin{eqnarray}
f_S=\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot \frac{h^\prime_S}{x_1
x_2 (M^{\prime2}-M^{\prime2}_0)}4(m_1^\prime x_2-m_2 x_1),
\label{eq:fS}
\end{eqnarray}
\begin{eqnarray}
f_{^3\!A}&=&-\frac{N_c}{4\pi^3M^\prime}\int dx_2 d^2p^\prime_\bot
\frac{h^\prime_{^3\!A}}{x_1 x_2 (M^{\prime2}-M^{\prime2}_0)}
\nonumber\\
&&\qquad\qquad\times\left[x_1 M^{\prime2}_0-m_1^\prime (m_1^\prime+m_2)-p^{\prime2}_\bot
-\frac{m_1^\prime-m_2}{w^\prime_{^3\!A}}\,p^{\prime2}_\bot \right],
\nonumber\\
f_{^1\!A}&=&\frac{N_c}{4\pi^3 M^\prime}\int dx_2 d^2p^\prime_\bot
\frac{h^\prime_{^1\!A}}{x_1 x_2 (M^{\prime2}-M^{\prime2}_0)}
\left(\frac{m_1^\prime-m_2}{w^\prime_{^1\!A}}\,p^{\prime2}_\bot \right),
\label{eq:fA}
\end{eqnarray}
where
\begin{eqnarray} \label{eq:vertex}
h^\prime_P&=&h^\prime_V
=(M^{\prime2}-M_0^{\prime2})\sqrt{\frac{x_1 x_2}{N_c}}
\frac{1}{\sqrt{2}\widetilde M^\prime_0}\varphi^\prime,
\nonumber\\
h^\prime_S &=&\sqrt{\frac{2}{3}}h^\prime_{^3\!A}
=(M^{\prime2}-M_0^{\prime2})\sqrt{\frac{x_1 x_2}{N_c}}
\frac{1}{\sqrt{2}\widetilde M^\prime_0}\frac{\widetilde
M^{\prime
2}_0}{2\sqrt{3}M^\prime_0}\varphi^\prime_p,
\nonumber\\
h^\prime_{^1\!A}&=& h^\prime_T =(M^{2\prime}-M_0^{\prime 2})\sqrt{\frac{x_1
x_2}{N_c}}\frac{1}{\sqrt{2}\widetilde M^\prime_0}\varphi'_p\, ,
\nonumber\\
w^\prime_V&=&M^\prime_0+m^\prime_1+m_2,\quad
w^\prime_{^3\!A}=\frac{\widetilde{M}'^2_0}{m^\prime_1-m_2},\quad
w^\prime_{^1\!A}=2\,,
\label{eq:h}
\end{eqnarray}
are the appropriate replacements of the vertex functions,
\begin{eqnarray}
H^\prime_M
&\to&\hat H^\prime_M
=H^\prime_M(\hat p^{\prime 2}_1,\hat p^2_2)
\equiv h^\prime_M,
\nonumber\\
W^\prime_M
&\to&\hat W^\prime_M
=W^\prime_M(\hat p^{\prime 2}_1,\hat p^2_2)
\equiv w^\prime_M,
\label{eq:contourA}
\end{eqnarray}
appearing in the matrix elements of annihilation of a meson state via weak currents,
and $\varphi'$ and $\varphi'_p$ are the light-front momentum
distribution amplitudes for $s$-wave and $p$-wave mesons,
respectively. There are several popular phenomenological
light-front wave functions that have been employed to describe
various hadronic structures in the literature. In the present
work, we shall use the Gaussian-type wave function \cite{Gauss}
\begin{eqnarray} \label{eq:Gauss}
\varphi^\prime
&=&\varphi^\prime(x_2,p^\prime_\perp)
=4 \left({\pi\over{\beta^{\prime2}}}\right)^{3\over{4}}
\sqrt{{dp^\prime_z\over{dx_2}}}~{\rm exp}
\left(-{p^{\prime2}_z+p^{\prime2}_\bot\over{2 \beta^{\prime2}}}\right),
\nonumber\\
\varphi^\prime_p
&=&\varphi^\prime_p(x_2,p^\prime_\perp)=\sqrt{2\over{\beta^{\prime2}}}~\varphi^\prime,\quad\qquad
\frac{dp^\prime_z}{dx_2}=\frac{e^\prime_1 e_2}{x_1 x_2 M^\prime_0}.
\label{eq:wavefn}
\end{eqnarray}
The parameter $\beta'$, which describes the momentum distribution,
is expected to be of order $\Lambda_{\rm QCD}$.
Note that with the explicit form of
$h'_P$ shown in Eq.~(\ref{eq:h}), the familiar expression of $f_P$
in the conventional light-front approach~\cite{Jaus91,Cheng97},
namely,
\begin{eqnarray} \label{eq:fP0}
f_P=2\frac{\sqrt{2N_c}}{16\pi^3}\int dx_2 d^2p^\prime_\bot \frac{1}{\sqrt{x_1
x_2} \widetilde M'_0}\,(m_1^\prime x_2+m_2
x_1)\,\varphi^\prime(x_2,p^\prime_\perp),
\end{eqnarray}
is reproduced. For decay constants of vector and axial-vector
mesons, we consider the case with the transverse polarization given by
\begin{eqnarray}
\varepsilon(\pm)=\left(\frac{2}{P^{\prime+}}\varepsilon_\bot\cdot
P^\prime_\bot,0,\varepsilon_\bot\right),\qquad
\varepsilon_\bot=\mp\frac{1}{\sqrt2}(1,\pm i).
\end{eqnarray}
For $m_1^\prime=m_2$, the meson wave function is symmetric with
respect to $x_1$ and $x_2$, and hence $f_S=0$, as it should be.
Similarly, it is clear that $f_{^1\!A}=0$ for $m^\prime_1=m_2$. The
SU(N)-flavor constraints on $f_S$ and $f_{^1\!A}$ are thus
satisfied.\footnote{We wish to stress that the vector decay constant obtained in the
conventional light-front model \cite{Jaus91} does not coincide
with the above result (\ref{eq:fV}) owing to the missing zero mode
contribution.}
\begin{table}[b!]
\caption{\label{tab:beta} The input parameter $\beta$ (in units of
GeV) in the Gaussian-type wave function (\ref{eq:wavefn}) for mesons. Note that
$\beta_{q\bar q}$ is used for the $(u\overline{u} + d\overline{d})/\sqrt{2}$ state.}
\begin{ruledtabular}
\begin{tabular}{|c|ccccc|}
$^{2S+1} L_J$
& $^1S_0$
& $^3S_1$
& $^3P_0$
& $^3P_1$
& $^1P_1$
\\
\hline
$\beta_{d\bar u}$
& $0.3077^{+0.0009}_{-0.0008}$
& $0.2815^{+0.0046}_{-0.0047}$
& $0.2983^{+0.0123}_{-0.0129}$
& $0.2983^{+0.0123}_{-0.0129}$
& $0.2983^{+0.0123}_{-0.0129}$
\\
$\beta_{q\bar q}$
& $0.3499^{+0.0136}_{-0.0129}$
& $0.2640^{+0.0031}_{-0.0032}$
& $0.2983\pm 0.0298$
& $0.2983\pm 0.0298$
& $0.2983\pm 0.0298$
\\
$\beta_{s\bar u}$
& $0.3479^{+0.0029}_{-0.0029}$
& $0.2926^{+0.0047}_{-0.0047}$
& $0.3224^{+0.0163}_{-0.0195}$
& $0.3224^{+0.0163}_{-0.0195}$
& $0.3224^{+0.0163}_{-0.0195}$
\\
$\beta_{s\bar s}$
& $0.3598^{+0.0220}_{-0.0208}$
& $0.3083\pm 0.0014$
& $0.3492\pm 0.0064$
& $0.3492\pm 0.0064$
& $0.3492\pm 0.0064$
\\
\hline
$\beta_{c\bar u}$
& $0.4656^{+0.0217}_{-0.0212}$
& $0.4255\pm 0.0426$
& $0.3890\pm 0.0389$
& $0.3890\pm 0.0389$
& $0.3890\pm 0.0389$
\\
$\beta_{c\bar s}$
& $0.5358^{+0.0137}_{-0.0135}$
& $0.4484\pm 0.0448$
& $0.3900\pm 0.0390$
& $0.3900\pm 0.0390$
& $0.3900\pm 0.0390$
\\
$\beta_{c\bar c}$
& $0.7690^{+0.0049}_{-0.0049}$
& $0.6492\pm 0.0069$
& $0.4200\pm 0.0420$
& $0.4200\pm 0.0420$
& $0.4200\pm 0.0420$
\\
\hline
$\beta_{b\bar u}$
& $0.5547^{+0.0260}_{-0.0261}$
& $0.5183\pm 0.0518$
& $0.5000\pm 0.0500$
& $0.5000\pm 0.0500$
& $0.5000\pm 0.0500$
\\
$\beta_{b\bar s}$
& $0.6103^{+0.0330}_{-0.0331}$
& $0.5589\pm 0.0559$
& $0.5500\pm 0.0550$
& $0.5500\pm 0.0550$
& $0.5500\pm 0.0550$
\\
$\beta_{b\bar c}$
& $0.9582\pm 0.0958$
& $0.8451\pm 0.0845$
& $0.6800\pm 0.0680$
& $0.6800\pm 0.0680$
& $0.6800\pm 0.0680$
\\
$\beta_{b\bar b}$
& $1.4514\pm 0.0132$
& $1.3267\pm 0.0100$
& $0.9993\pm 0.0999$
& $0.9993\pm 0.0999$
& $0.9993\pm 0.0999$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
To perform numerical computations of decay constants and form factors, we need to
specify the input parameters in the covariant light front model.
These are the constituent quark masses and the shape parameter $\beta$
appearing in the Gaussian-type wave function (\ref{eq:wavefn}).
For constituent quark masses, we
use~\cite{Jaus96,Cheng97,Hwang02,Jaus99, Cheng04}
\begin{eqnarray} \label{eq:quarkmass}
m_{u,d}=0.26\,{\rm GeV},\qquad m_s=0.45\,{\rm GeV},\qquad
m_c=1.40\,{\rm GeV},\qquad m_b=4.64\,{\rm GeV}.
\end{eqnarray}
Shown in Tables~\ref{tab:beta} and \ref{tab:f} are the input parameter
$\beta$ and decay constants, respectively. In Table \ref{tab:f}
the decay constants in parentheses are used to determine $\beta$ using the
analytic expressions in the covariant light-front model as given above.
For most of $s$-wave mesons, and a few axial vector mesons, these are fixed from
the latest decay rates given in the Particle Data Group ~\cite{PDG10},
or other analysis based on some experimental results. For decay constants of
some heavy flavor mesons, we have used recent lattice results to
fix $\beta$. For the remaining $p$-wave mesons, we use the $\beta$ parameters
obtained in the ISGW2 model~\cite{ISGW2}, the improved version of the ISGW
model, up to some simple scaling. In this paper, we have investigated
the variation of the form factors and their slope parameters for $q^2$
dependence with the variation of $\beta$ values.
Wherever the experimental information is available, we have
used that to fix the errors for the corresponding $\beta$ values,
otherwise arbitrarily introduced an uncertainty of $10 \%$ in $\beta$
for some $s$-wave and $p$-wave mesons.
\begin{table}[t!]
\caption{\label{tab:f} Meson decay constants (in units of MeV)
obtained by using Eqs.~(\ref{eq:fP}), (\ref{eq:fS}), (\ref{eq:fV})
and (\ref{eq:fA}). Those in parentheses are taken as inputs to
determine the corresponding $\beta$'s shown in
Table~\ref{tab:beta}. Decay constants of some $p$-wave mesons are
also used as inputs (see the text for details). Here
$f_{q\bar q}$ denotes decay constant for the
$(u\overline{u} + d\overline{d})/\sqrt{2}$ state.}
\begin{ruledtabular}
\begin{tabular}{|c|ccccc|}
$^{2S+1} L_J$
& $^1S_0$
& $^3S_1$
& $^3P_0$
& $^3P_1$
& $^1P_1$
\\
\hline
$f_{d\bar u}$
& $(130.41\pm 0.20)$
& $(215\pm 5)$
& $0$
& $(-203\mp 18)$
& $0$
\\
$f_{q\bar q}$
& $(139.54\pm 2.62)$
& $(195\pm 3)$
& $0$
& $-193^{-43}_{+38}$
& $0$
\\
$f_{s\bar u}$
& $(156.1\pm 0.9)$
& $(217\pm 5)$
& $34.9^{+1.4}_{-1.8}$
& $(-212^{-23}_{+26})$
& $20.4^{+1.5}_{-1.8}$
\\
$f_{s\bar s}$
& $(174.75\pm7.83)$
& $(228\pm 2)$
& $0$
& $(-230\mp9)$
& $0$
\\
\hline
$f_{c\bar u}$
& $(206.7\pm8.9)$
& $(245)^{+35}_{-34}$
& $107\pm13$
& $-177^{-38}_{+34}$
& $59.6^{+9.8}_{-9.5}$
\\
$f_{c\bar s}$
& $(254.6\pm5.9)$
& $(272)^{+39}_{-38}$
& $74.4^{+10.4}_{-10.6}$
& $-159^{-36}_{+32}$
& $42.2^{+7.6}_{-7.3}$
\\
$f_{c\bar c}$
& $(394.7\pm2.4)$
& $(411\pm 6)$
& $0$
& $-105^{-26}_{+23}$
& $0$
\\
\hline
$f_{b\bar u}$
& $(193\pm11)$
& $(196)^{+28}_{-27}$
& $143\pm21$
& $-155^{-30}_{+28}$
& $83.6^{+14.3}_{-13.6}$
\\
$f_{b\bar s}$
& $(231\pm15)$
& $(229)^{+32}_{-31}$
& $139\pm22$
& $-166^{-34}_{+31}$
& $82.6^{+15.0}_{-14.2}$
\\
$f_{b\bar c}$
& $440^{+51}_{-52}$
& $440^{+51}_{-52}$
& $90.6^{+17.5}_{-16.6}$
& $-155^{-37}_{+33}$
& $52.0^{+11.2}_{-10.3}$
\\
$f_{b\bar b}$
& $(708\pm8)$
& $(708\pm8)$
& $0$
& $-185^{-48}_{+42}$
& $0$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
Several remarks are in order:
(i) Decay constants of the charged pseudoscalar mesons, $\pi^+, K^+, D^+, D^+_s,$ and $B^-$
(and their charge-conjugate partners) can be determined from their purely leptonic
decay rates. These mesons formed from a quark and anti-quark can decay to a charged lepton pair
when their constituents annihilate via a virtual $W$ boson. Now quite precise measurements
are available for the branching fractions of $P \to \ell \nu_{\ell}$ decays \cite{PDG10}. Following the
analysis of Rosner and Stone \cite{Rosner10} for the available
branching fractions, we take
$f_{\pi}= 130.41\pm 0.20, f_K = 156.10\pm0.85, f_D = 206.7\pm8.9$
(all in MeV) to fix the $\beta$ parameters of the respective mesons.
(ii) For fixing $\beta_{D_s}$, we have taken the world average value $254.6\pm5.9$ MeV
for $f_{D_s}$ given by the Heavy Flavor Averaging Group \cite{HFAG10} based on
the BaBar, Belle and CLEO measurements of ${\cal B}( D^+_s \rightarrow \mu^+\nu)$
and ${\cal B}(D^+_s \rightarrow \tau^+\nu)$. This value can be compared well to the
results from the two precise lattice QCD calculations
$f_{D_s} = 248.0 \pm 2.5 $ MeV and $249 \pm 11$ MeV, respectively, from the HPQCD
Collaboration \cite{HPQCD10} and the Fermilab/MILC Collaboration \cite{MILC08}.
For the bottom sector, the Belle and BaBar collaborations have found
evidence for $B^- \rightarrow \tau^- \nu$ decay in $e^+ e^-\rightarrow
B^-B^+$ collisions at the $\Upsilon(4S)$ energy, however, the errors are rather
large in the measured branching fractions with the computed average value
${\cal B}(B^- \rightarrow \tau^- \nu) = 1.72^{+0.43}_{-0.42} \times 10^{-4}$. Further a more
accurate value of $|V_{ub}|$ is required for the determination of $f_B$.
Considering the large uncertainties on $V_{ub}$ and the branching fraction
measurements for $B^- \rightarrow \tau^- \nu$, and sensitivity of this decay
to the new physics, we rely upon $f_B = 193 \pm 11 $ MeV, used in \cite{Rosner10} as the
average of the two lattice results $f_B = 195 \pm 11$ MeV \cite{MILC08} and
$f_B = 190 \pm 13 $ MeV \cite{HPQCD09}, to fix the input parameter $\beta_{B}$.
Likewise, for $B_s$ meson, we use the lattice prediction of $f_{B_s} = 231 \pm 15$ MeV
\cite{HPQCD09} for determining $\beta_{B_s}$.
(iii) The decay constants of the diagonal pseudoscalar mesons
$\pi^0, \eta, \eta^\prime$ and $\eta_c$,
in principle, could be obtained from $P \rightarrow \gamma \gamma$ branching fractions.
In the case of $\pi^0$, the value of $f_{\pi^0}= 130 \pm 5$ MeV \cite{Suzuki03} has been extracted
from the measured $\pi^0 \rightarrow \gamma \gamma$ decay width, which is compatible
with $f_{\pi^{\pm}}$, as is expected from isospin symmetry. However, decay constants
of the $\eta -\eta^\prime$ system cannot be extracted from two-photon decay rates
alone and get more complicated due to the $\eta -\eta^\prime$ mixing, the chiral
anomaly and gluonium mixing \cite{Feldmann00, ChengLi09}. For describing the
mixing between $\eta$ and $\eta^\prime$, it is more convenient to employ
the flavor states $(u\overline{u} + d\overline{d})/\surd {2}$, and
$ (s\overline{s})$ labeled by the $\eta_q$ and $\eta_s$, respectively. We then write
\begin{eqnarray} \label{eq:EtaMixing}
\eta = \eta_q \cos\phi - \eta_s \sin\phi,
\nonumber\\
\eta^\prime = \eta_q \sin\phi + \eta_s\cos\phi,
\end{eqnarray}
where $\phi = (39.3 \pm 1.0)^ \circ$ follows from the analysis of Feldmann {\it et al.} \cite{Feldmann00}
to fit the experimental data. This analysis also gives
$f_{\eta}/f_{\pi}= 1.07 \pm 0.02$ and $f_{\eta^{\prime}}/f_{\pi}= 1.34 \pm 0.06$, which are used
in the present work.
For $\eta_c$, the decay width is poorly known with PDG \cite{PDG10} estimate
given as $\Gamma (\eta_c \rightarrow \gamma \gamma)=7.2\pm2.1$ keV
giving $f_{\eta_c} = 0.4 \pm 0.1 $ GeV. Alternatively,
one may extract $f_{\eta_c}$ from $B \rightarrow \eta_c K $ decay using the
factorization approximation, for which CLEO \cite{Edwards01} obtained $f_{\eta_c} = 335 \pm 75$ MeV.
In the literature, $f_{\eta_c}$ is expected to be quite close to $f_{J/\psi}$
on the basis of quark model considerations \cite{Gourdin95}. Recently, the
HPQCD collaboration \cite{HPQCD10} has reported a more precise result for
$f_{\eta_c}$ to be $394.7\pm2.4$ MeV consistent with other estimates,
and is in fact very close to the experimental result $f_{J/\psi}= 410.6\pm 6.2$ MeV
obtained from the leptonic decay width of $J/\psi$ \cite{PDG10}. So we use the
lattice prediction to fix $\beta_{\eta_c}$. In the absence of
any experimental estimate for $f_{\eta_b}$, we shall assume
$f_{\eta_b} \approx f_{\Upsilon}$ to fix $\beta_{\eta_b}$ following the heavy-quark
spin symmetry.
(iv) For vector mesons, we extract the decay constants for diagonal states from
the experimental values of their respective branching fractions of leptonic decays
$V \rightarrow l^+ l^-$ decays \cite{PDG10}. Thus we obtain
$f_{\rho^0}= 221.20 \pm 0.94,~ f_{\omega} = 194.60 \pm 3.24,~
f_{\phi}= 227.9 \pm 1.5,~ f_{J/\psi}= 410.6\pm 6.2$ and $f_{\Upsilon} = 708.0\pm7.8$ (all in MeV)
for ideal mixing, and use them to fix the $\beta_V$ parameters of the respective mesons.
(v) The decay constant $f_V$ determines not only the coupling of the neutral
vector mesons to a photon, but also the coupling of charged vector mesons, like
$\rho^{\pm}$ and $K^{*\pm}$,
to the weak vector bosons $W^{\pm}$. There are no data available for the leptonic
decay of these charged vector mesons, but the couplings can be extracted indirectly
from the decays $\tau \rightarrow \rho \nu_{\tau}$ and
$\tau \rightarrow K^* \nu_{\tau}$. With the experimental values for the
branching fractions of these decays ${\cal B}(\tau \rightarrow \rho \nu_{\tau})= 25.02 \% $
and ${\cal B}(\tau \rightarrow K^* \nu_{\tau})= 1.28 \% $, the decay width formula
\begin{eqnarray} \label{eq:tauVdec}
\Gamma(\tau \rightarrow V \nu_{\tau}) = {{G^2_F} \over {16 \pi}} |V_{q_1 q_2}|^2 f^2_{V} {{(m^2_{\tau}+2 m^2_V)(m^2_{\tau}- m^2_V)^2} \over {m^3_{\tau}}},
\end{eqnarray}
where $V_{q_1 q_2}$ is the appropriate CKM- factor corresponding to the vector meson $V$,
yield $f_{\rho^{\pm}} = 209\pm4$ MeV and $f_{K^{* \pm}} = 217 \pm 5$ MeV, respectively.
It is worth noting that the difference in $f_{\rho^0}$ and $f_{\rho^{\pm}}$
seems consistent with the expected size of isospin breaking, and we take the
average of the two values, i.e., $f_{\rho}= 215\pm 5$ MeV, the error chosen
so as to satisfy the two cases in extreme limits.
(vi) Contrary to the non-strange charmed
meson case, where $D^*$ has a slightly larger decay constant than
$D$, the recent measurements of $B\to D_s^{(*)}D^{(*)}$
\cite{PDG10,BaBarDs} indicate that the decay constants of $D_s^*$
and $D_s$ are relatively similar. As for the decay constant of $B^*$, a recent lattice calculation
yields $f_{B^*}/f_B=1.01\pm0.01^{+0.04}_{-0.01}$ \cite{Bernard}.
Explicitly, for naked charmed and bottom states $D^*, D^*_s, B^*$, and $B^*_s$,
we have used the lattice predictions, $f_{D^*}= 245, f_{D^*_s}= 272, f_{B^*}= 196$, and $f_{B^*_s} = 229$
(all in MeV) \cite{Becirevic99} to fix the central value of the respective parameters $\beta$,
and allow $10\%$ variation in each case, giving decay constant ratios as
$f_{D^*}/f_D = 1.18 \pm 0.17, f_{D^*_s}/f_{D_s} = 1.07 \pm 0.15$, and
$ f_{B^*}/f_B \approx f_{B^*_s}/f_{B_s} \approx 1.0 \pm 0.15 $, to leave the
scope for matching with other results.
(vii) For axial vector mesons, there are two different nonets of
$J^P = 1^+$ in the quark model as the orbital excitation of the
$q \overline{q}$ system. In terms of the spectroscopic notation
$^{2S+1} L_J$, there are two types of $p$-wave axial vector mesons,
namely, $^3 P_1$ and $^1 P_1$, which have distinctive C quantum
numbers, $C=+$ and $C=-$, respectively. Experimentally,
the $J^{PC}= 1^{++}$ nonet consists of $a_1 (1260),
f_1 (1285), f_1 (1420)$, and $K_{1A}$, while the $J^{PC}= 1^{+-}$ nonet has
$b_1 (1235), h_1 (1170), h_1 (1380)$, and $K_{1B}$.
(viii) It is generally argued that $a_1(1260)$ should have a
similar decay constant as the $\rho$ meson.
Presumably, $f_{a_1}$ can be extracted from the decay
$\tau\to a_1(1260)\nu_\tau$. Though this decay is not shown in the
Particle Data Group \cite{PDG10}, an experimental value of
$|f_{a_1}|=203\pm 18$ MeV is nevertheless quoted in
\cite{Bloch}.\footnote{The decay constant of $a_1$ can be tested
in the decay $B^+\to \bar D^0a_1^+$ which receives the main
contribution from the color-allowed amplitude proportional to
$f_{a_1}F^{BD}(m_{a_1}^2)$.} The $a_1 (1260)$ decay constant
$f_{a_1} = 238 \pm 10$ MeV obtained using the QCD sum rule method
\cite{Yang07} is slightly higher than this value as well
as $f_{\rho}= 215$ MeV. In Table \ref{tab:f} we have employed $f_{a_1}=-203\pm 18$ MeV
as input following our sign convention.
(ix) The nonstrange axial-vector mesons, for example, $a_1 (1260)$
and $b_1 (1235)$ cannot have mixing because
of the opposite $C$-parities. On the contrary, physical strange
axial-vector mesons are the mixture of $^3P_1$ and $^1P_1$ states, while
the heavy axial-vector resonances are generally taken as the mixture of $P_1^{1/2}$ and
$P_1^{3/2}$. For example, the physical
mass eigenstates $K_1 (1270)$ and $K_1 (1400)$ are a mixture of
$K_{1A}$ and $K_{1B}$ states owing to the mass difference of the
strange and nonstrange light quarks:
\begin{eqnarray} \label{eq:K1mixing}
K_1(1270)=K_{1A} \sin\theta_{K_1} +K_{1B}\cos\theta_{K_1},
\nonumber\\
K_1(1400)=K_{1A} \cos\theta_{K_1}-K_{1B}\sin\theta_{K_1}.
\end{eqnarray}
Using the experimental results ${\cal B}(\tau \rightarrow K_1 (1270) \nu_{\tau})=
(4.7 \pm 1.1) \times 10^{-3} $ and $\Gamma(\tau \rightarrow K_1 (1270) \nu_{\tau})/
[ \Gamma (\tau \rightarrow K_1 (1270) \nu_{\tau})+\Gamma (\tau \rightarrow K_1 (1400)
\nu_{\tau})]= 0.69\pm 0.15$ \cite{PDG10}, and the decay width formula similar
to that given in Eq. (\ref{eq:tauVdec}) with the replacement $V \to A$,
we obtain \footnote{The large experimental error with the $K_1 (1400)$
production in the $\tau$ decays, namely ${\cal B}(\tau \rightarrow K_1 (1400) \nu_{\tau})=
(1.7 \pm 2.6) \times 10^{-3}$ \cite{PDG10}, does not provide sensible
information for the $K_1 (1400)$ decay constant}
\begin{eqnarray} \label{eq:K1decayConst}
| f_{K_1(1270)}|= 169.5^{+18.8}_{-21.2}\, \textrm{MeV},
\nonumber\\
| f_{K_1(1400)}|= 139.2^{+41.3}_{-45.6}\, \textrm{MeV}.
\end{eqnarray}
These decay constants are related to $f_{K_{1A}}$ and $f_{K_{1B}}$ through
\begin{eqnarray} \label{eq:K1decayConstMixing}
m_{K_1 (1270)} f_{K_1 (1270)} = m_{K_{1A}} f_{K_{1A}} \sin\theta_{K_1} + m_{K_{1B}} f_{K_{1B}} \cos\theta_{K_1},
\nonumber\\
m_{K_1 (1400)} f_{K_1 (1400)} = m_{K_{1A}} f_{K_{1A}} \cos\theta_{K_1} - m_{K_{1B}} f_{K_{1B}} \sin\theta_{K_1},
\end{eqnarray}
where use of Eq. (\ref{eq:K1mixing}) and expressions for decay constants have been made.
From the analytic expressions of decay constants given in Eq. (\ref{eq:AM}),
it is clear that $m_{K_{1A}} f_{K_{1A}}$ and $m_{K_{1B}} f_{K_{1B}}$ are functions
of $\beta_{K_1}$ and quark masses only. In other words, they do not depend on
$m_{K_{1A, 1B}}$ and hence $\theta_{K_1}$. Eq. (\ref{eq:K1decayConstMixing})
leads to the following relation:
\begin{eqnarray} \label{eq:K1MassDecayConst}
m^2_{K_1 (1270)} f^2_{K_1 (1270)} + m^2_{K_1 (1400)} f^2_{K_1 (1400)} =
m^2_{K_{1A}} f^2_{K_{1A}} + m^2_{K_{1B}} f^2_{K_{1B}}.
\end{eqnarray}
This relation, being independent of the mixing angle $\theta_{K_1}$, has been
used \cite{Cheng10} to determine the central value of the parameter
$\beta_{K_1}$ to be 0.3224 GeV. However, to calculate the individual decay constants,
masses of $K_1$ mesons are needed for which the mixing angle $\theta_{K_1}$ is required.
From Eq. (\ref{eq:K1mixing}), the masses of the $K_{1A}$ and $K_{1B}$ can be expresses as
\begin{eqnarray} \label{eq:K1masses}
m^2_{K_{1A}} = m^2_{K_1 (1270)} \sin^2\theta_{K_1} + m^2_{K_1 (1400)} \cos^2\theta_{K_1},
\nonumber\\
m^2_{K_{1B}} = m^2_{K_1 (1270)} \cos^2\theta_{K_1} + m^2_{K_1 (1400)} \sin^2\theta_{K_1}.
\end{eqnarray}
There exists several estimations on the mixing angle in the literature
\cite{Suzuki, Burakovsky, Cheng03} differing in the value and sign convention.
These often employ masses, partial decay rates of $K_1 (1270)$ and $K_1 (1400)$,
and $\tau$ decay rates to these mesons.\footnote{The relative signs of the decay
constants, form factors, and mixing angles of the axial vector mesons were often
confusing in the literature. The sign of mixing angle is intimately related to the
relative sign of the $K_{1A}$ and $K_{1B}$ states. For a detailed discussion, refer
to \cite{Cheng07, Cheng10, Cheng08}.} Note that in the CLF quark model, the sign
of $f_{K_{1A}}$ is negative, whereas $f_{K_{1B}}$ is positive. With this
sign convention, from Eq. (\ref{eq:K1decayConstMixing}), the following two
solutions \cite{Cheng10} have been obtained:
\begin{eqnarray} \label{eq:K1mixingValue}
\theta_{K_1} = \left\{ \begin{array}{c}
+50.8^\circ ~ \textrm{solution~ I}, \\
-44.8^\circ ~ \textrm{solution ~II}.
\end{array} \right.
\end{eqnarray}
The second solution is ruled out by the experimental data for $ B \rightarrow K_1 (1270)/K_1 (1400)
+ \gamma$ decays \cite{Cheng10}. For $\theta_{K_1} = 50.8^\circ$, masses of $^3P_1$ and $^1P_1$
states come out to be,
\begin{eqnarray} \label{eq:K1ABmasses}
m_{K_{1A}} = 1.26 ~\textrm{GeV}, ~~ m_{K_{1B}} = 1.52~\textrm{ GeV},
\end{eqnarray}
corresponding to
\begin{eqnarray} \label{eq:K1decConstPhy}
f_{K_1(1270)}= -170~ \textrm{MeV}, ~~ f_{K_1(1400)} = 139~ \textrm{MeV}.
\end{eqnarray}
Note that obtained value for $\beta_{K_1}$ lies between $\beta_{K}$ and $\beta_{K^*}$.
(x) Like $s$-wave mesons, there are mixing between singlet
and octet states of $p$-wave mesons also, equivalently, between $u\overline{u} + d \overline{d}$ and
$s \overline{s}$ components. For axial vector states $f_1 (1285)$ and $f_1 (1420)$,
the mixing can be written as
\begin{eqnarray} \label{eq:f1mixing}
f_1(1285)=f_{1q} \sin\alpha_{f_1} + f_{1s}\cos\alpha_{f_1},
\nonumber\\
f_1(1420)=f_{1q} \cos\alpha_{f_1} - f_{1s}\sin\alpha_{f_1},
\end{eqnarray}
where
$f_{1q}= (u\overline{u} + d\overline{d})/\surd {2}$, and
$f_{1s}$ is pure $(s\overline{s})$ state. The mixing angle $\alpha_{f_1} $
is related to the singlet-octet mixing angle
$\theta_{f_1}$ by the $ \alpha_{f_1} = \theta_{f_1} + 54.7^\circ$,
where the latter mixing angle is defined by
\begin{eqnarray} \label{eq:f1mixing18}
f_1(1285)= f_1 \cos\theta_{f_1} + f_8 \sin\theta_{f_1} ,
\nonumber\\
f_1(1420)= - f_1 \sin\theta_{f_1} + f_8 \cos\theta_{f_1} .
\end{eqnarray}
The magnitude of the angle is given by the mass relations
\begin{eqnarray} \label{eq:f1mixingAngle}
\tan^2 \theta_{f_1} ={ { 4 m^2_{K_{1A}} - m^2_{a_1} - 3 m^2_{f_1 (1420)} } \over
{- 4 m^2_{K_{1A}} + m^2_{a_1} + 3 m^2_{f_1 (1285)}} },
\end{eqnarray}
while the sign of the angle can be determined from
\begin{eqnarray} \label{eq:f1mixingAngle1}
\tan \theta_{f_1} ={ { 4 m^2_{K_{1A}} - m^2_{a_1} - 3 m^2_{f_1 (1420)}} \over
{2 \surd {2}( m^2_{a_1} - m^2_{K_{1A}}) } }.
\end{eqnarray}
We thus obtain $\alpha_{f_1} = 94.9^\circ$, i.e., $\theta_{f_1} = 40.2^\circ$.
Denoting the mass of the $f_{1s}$ component as $m_{f_{1s}}$, we have
\begin{eqnarray} \label{eq:f1ssmass}
m^2_{f_{1s}} = m^2_{f_1 (1420)} \sin^2 \alpha_{f_1} + m^2_{f_1 (1285)} \cos^2 \alpha_{f_1},
\end{eqnarray}
which yields $m_{f_{1s}}= 1.425$ GeV. Using the mixing angle and $m_{s \overline{s}}$,
the decay constant $f_{f_{1s}}$ of the $^3 P_1$ axial vector meson with a pure
$s \overline{s}$ quark content has been determined to be $-230\pm 9 $ MeV \cite{Yang10}.
Consequently, $\beta_{f_{1s} }$ gets fixed in the present CLF model to
be $0.3492\pm0.0064$ \cite{Cheng10}. For the purpose
of an estimation, for the remaining axial vector mesons, we use the $\beta$ parameters
obtained in the ISGW2 model~\cite{ISGW2} up to some simple scaling, and $10 \%$ uncertainty
has been assigned to them arbitrarily to study its effects on their decay constants and
the corresponding form factors.
(xi) The $\beta$ values are kept same for other
$p$-wave mesons, scalar $(J^{PC}=0^{++})$ and axial vector
$(J^{PC}=1^{+-})$ mesons, as that of the $(J^{PC}=1^{++})$ mesons
having the same flavor quantum numbers.
So their decay constants are calculated respectively as shown in Table II.
The $\beta$ parameters for $p$-wave states of the charmed and bottom states
are smaller when compared to the respective $\beta_{P,V}$ values.
(xii) Situation regarding the decay constant for the $^1 P_1$ mesons is different
from the $^3 P_1$ mesons. First of all, its decay
constant vanishes in the isospin or SU(3) limit. In fact, because of charge
conjugation invariance, the decay constant of the nonstrange neutral
meson $b^0_1 (1235)$ must be zero. In the isospin limit, the decay constant
of the charged $b_1$ vanishes due to the fact that the $b_1$ has even
$G$-parity and that the relevant weak axial-vector current is odd under
$G$ transformation. Hence, $f_{b^+_1 (1235)}$ is very small in reality, arising due to the
small mass difference between $u$ and $d$ quark masses. In the present covariant
light-front quark model, if we increase the constituent $d$ quark mass by
an amount of $5\pm2$ MeV relative to the $u$ quark one, we find
$f_{b^+_1 (1235)}= 0.6 \pm 0.2 $ MeV which is highly suppressed.
\footnote {In \cite{Laporta06}, the decay constants
of $a_1$ and $b_1$ are derived using the $K_{1A} -K_{1B}$ mixing angle
$\theta_{K_1}$ and SU(3) symmetry to be $(f_{b_1}; f_{a_1}) = (74; 215)$ MeV
for $\theta_{K_1}= 32^\circ$ and $(-28; 223)$ MeV for $\theta_{K_1}= 58^\circ$.
It seems to us that the magnitude of $b_1$ decay constant derived
in this manner is too big.}
Similar to the $f_1(1285) -f_1 (1420)$ mixing in the
$J^{PC} = 1^{++}$ nonet, the $h_1 (1170)- h_1 (1380)$
mixing can be described for $J^{PC} = 1^{+-}$ with the replacement
$ f_1 (1285)\rightarrow h_1 (1170), f_1 (1420)\rightarrow h_1 (1380)$, and
$\theta_{f_1} \rightarrow \theta_{h_1}, \alpha_{f_1} \rightarrow \alpha_{h_1}$,
leading to $\alpha_{h_1} = 54.7^\circ$ \cite{Cheng10}, i.e., $\theta_{h_1} = 0^\circ$.
However, like $b^0_1$, decay constants of these mesons also vanish.
In fact, in the SU(3) limit, $f_{^1 P_1} = 0$ should follow for strange mesons also
in this nonet. However, for the strange mesons $K_{1B}$, nonzero decay constant
would arise through SU(3) breaking, and we obtain $f_{K_{1B}}= 20.4^{+2.5}_{-1.8}$ MeV.
For charmed and bottom axial vector mesons, the results given in Table \ref{tab:f} translate
to $f^{1/2}_{D_1}= 179^{+37}_{-34}, f^{3/2}_{D_1}=-53.6^{-14.0}_{+12.1},$
$f^{1/2}_{D_{1s}}= 154^{+34}_{-31}, f^{3/2}_{D_{1s}}=-57.3^{-14.7}_{+12.8},$
$f^{1/2}_{B_1}= 175^{+33}_{-30}, f^{3/2}_{B_1}=-21.4^{-5.7}_{+4.9},$
$f^{1/2}_{B_{1s}}= 183^{+37}_{-34}, f^{3/2}_{B_{1s}}=-28.3^{-7.5}_{+6.4},$
$f^{1/2}_{B_{1c}}= 157^{+37}_{-33}$, and $ f^{3/2}_{B_{1c}}=-47.3^{-12.4}_{+10.7}$
(all in MeV). The errors shown here occur due to the $10 \%$ arbitrary uncertainty
assigned to their $\beta$ values. Note that the decay constants of $^3P_1$
and $P_1^{3/2}$ states have opposite
signs to that of $^1P_1$ or $P_1^{1/2}$ as can be easily seen from
Eq. (\ref{eq:Phalf}).
(xiii) Similarly for scalar mesons, their decay constants also vanish
in the SU(N) limit, as has been shown above Eq.(\ref{eq:Seom}) by applying equations
of motion. However, due to SU(N) breaking, only off-diagonal scalar
mesons can have nonzero decay constants, which have been given in
Table \ref{tab:f}. In the present covariant
light-front quark model, if the constituent $d$ quark mass is increased by
an amount of $5\pm2$ MeV relative to the $u$ quark one, we find
$|f_{a^{\pm}_0 (1450)}|= 1.1 \pm 0.4 $ MeV which is highly suppressed, whereas
SU(3) breaking yields $|f_{K^{*\pm}_0}|\approx 35$ MeV. Thus it
is clear that the decay constant of light scalar resonances remain largely suppressed
relative to that of the pseudoscalar mesons owing to the small mass difference
between the constituent quark masses, though this suppression
becomes less restrictive for heavy scalar mesons because of heavy
and light quark mass imbalance, and decay constants are of the order of hundred MeV.
Note that what is the underlying quark structure of light scalar
resonances is still controversial. While it has been widely
advocated that the light scalar nonet formed by $\sigma(600)$,
$\kappa(800)$, $f_0(980)$ and $a_0(980)$ can be identified
primarily as four-quark states, it is generally believed that the
nonet states $a_0(1450)$, $K_0^*(1430)$, $f_0(1370)$ and
$f_0(1500)/f_0(1710)$ are the conventional $q\bar q'$ states \cite{Close}. Therefore, the prediction of
$f_{K^*_0} \approx 35 $ MeV for the scalar meson in the $s \bar u$ content (see
Table \ref{tab:f}) is most likely designated for the $K_0^*(1430)$
state. Notice that this prediction is slightly smaller than the
result of $42$ MeV obtained in \cite{Maltman} based on the
finite-energy sum rules, and far less than the estimate of
$(70\pm10)$ MeV in \cite{Chernyak}. It is worth remarking that
even if the light scalar mesons are made from $4$ quarks, the decay
constants of the neutral scalars $\sigma(600)$, $f_0(980)$ and
$a_0^0(980)$ must vanish owing to charge conjugation invariance.
(xiv) In this work, we have only considered
the scalar nonet with masses above 1 GeV, for which the
quark content of $a_0(1450)$ and $K_0^*(1430)$ is quite
obvious, whereas the internal structure of the isoscalars $f_0(1370)$,
$f_0(1500)$ and $f_0(1710)$ in the same nonet is controversial
and less clear. Since not all the three isosinglet scalars can be
accommodated in the $q\bar q$ nonet picture, one of them should be
primarily a scalar glueball . Among them, it has been quite
controversial as to which of these
is the dominant scalar glueball. It has been
advocated that $f_0(1710)$ is mainly $(s\overline{s})$ and $f_0(1500)$ mostly
gluonic \cite{Close1, Amsler10}. However, this scenario encounters several
insurmountable difficulties, see \cite{Kleefeld, HYCScalar06} for detailed
discussions. Based on two simple and robust results as the input
for the mass matrix, the analysis in \cite{HYCScalar06} shows that in the limit of
exact SU(3) symmetry, $f_0(1500)$ is an SU(3) isosinglet octet
state and is degenerate with $a_0(1450)$. In the absence of glueball-quarkonium mixing, $f_0(1370)$
becomes a pure SU(3) singlet and $f_0(1710)$
the pure glueball. When the
glueball-quarkonium mixing is turned on, there will be some
mixing between the glueball ($G$) and the SU(3)-singlet $q\bar{q}$.
The mixing matrix obtained in this model has
the form \cite{HYCScalar06}:
\begin{eqnarray} \label{eq:wf}
\left(\begin{matrix} { f_0(1370) \cr f_0(1500) \cr f_0(1710) \cr} \end{matrix}\right)=
\left( \begin{matrix} { 0.78 & 0.51 & -0.36 \cr
-0.54 & 0.84 & 0.03 \cr
0.32 & 0.18 & 0.93 \cr}
\end{matrix}\right)\left(\begin{matrix}{f_{0q} \cr
f_{0s} \cr G \cr}\end{matrix}\right),
\end{eqnarray}
where $f_{0q}= (u\overline{u} + d\overline{d})/\surd {2}$, and
$f_{0s}$ is pure $(s\overline{s})$ state,
with masses $1.474$ GeV and $1.5$ GeV, respectively.
It is evident that $f_0(1710)$ is composed primarily of the scalar
glueball, $f_0(1500)$ is close to an SU(3) octet, and $f_0(1370)$
consists of an approximate SU(3) singlet with some glueball
component ($\sim 10\%$). Note that the recent quenched and unquenched lattice
calculations all favor a scalar glueball
mass close to 1700 MeV \cite{ChenMorningstar}.
(xv) In principle, the decay constant of the scalar strange charmed
meson $D^*_{s0}$ can be determined from the hadronic decay $B\to
\overline DD_{s0}^*$ since it proceeds only via external $W$-emission.
Indeed, a measurement of the $D\bar D_{s0}^*$ production in
$B$ decays by Belle \cite{BelleDs-1} indicates a $f_{D_{s0}^*}$ of
order 60 MeV \cite{Cheng:2003id} which is close to the calculated value
of 71 MeV (see Table \ref{tab:f}).
In our earlier work \cite{Cheng04}, we have discussed
more about $\overline DD_s^{**}$ productions in $B$ decays. The
smallness of the decay constant $f_{D_{s0}^*}$ relative to
$f_{D_s}$ can be seen from Eqs. (\ref{eq:fP}) and (\ref{eq:fS})
that
\begin{eqnarray}
f_{D_s(D_{s0}^*)}\propto \int dx_2\cdots[m_c x_2\pm m_s(1-x_2)].
\end{eqnarray}
Since the momentum fraction $x_2$ of the strange quark in the
$D_s(D_{s0}^*)$ meson is small, its effect being constructive in
$D_s$ case and destructive in $D_{s0}^*$ is sizable and explains
why $f_{D_{s0}^*}/f_{D_s}\sim 0.2$.
(xvi) For $D$ and $B$ systems, it is clear from Table~\ref{tab:f} that
$|f_{A^{3/2}}|\ll f_S < f_{A^{1/2}}$, in accordance with the
expectation from HQS [cf. Eq. (\ref{eq:HQSf})]. Decay constants of
$p$-wave charmed and bottom mesons
have been obtained using the Bethe-Salpeter method \cite{Wang07},
which are consistent with values obtained in Table II
for the bottom sector and $^1 P_1$ charmed mesons,
and are slightly higher than that of other $p$-wave charmed
mesons. However, our values for $D^*_{s0}$ and $D_{s1}$ match well with the results
$f_{D^*_{s0}}= 67.1 \pm 4.5 $ MeV and $f_{D_{s1}}= 144.5 \pm 11.1 $ MeV
obtained in \cite{Faessler07} based on the analysis of
$B\rightarrow D^* + D^*_{s0}/D_{s1}$ decays. For charmed $(c\overline{u})$
meson, our estimate $f_{D^{1/2}_1} = 179^{+37}_{-34} $ MeV is consistent with values
$f_{D^{1/2}_1}= 196 \pm93 $ MeV and $206 \pm120 $ MeV obtained in \cite{Jugeau05}
from the analysis of $\overline{B} \rightarrow D^{1/2}_1 \pi $ and
$\overline{B} \rightarrow D^{1/2}_0 \pi $ decays, respectively.
(xvii) The $\beta$ values used in the present analysis often differ from the
ones given in the earlier work ~\cite{Cheng04} to match with the
decay constants based on the latest data. For the same reason,
the strange quark mass $m_s = 0.45 $ GeV used here is different from
the values $0.37$ GeV used earlier ~\cite{Cheng04}. This choice of the strange
quark mass has to be made to obtain $f_{K^*_0} = 0.35$ MeV
based on the analysis of the $B$ decays emitting $K^*_0$
\cite{Cheng06,ChengChua10}. Otherwise, for the strange
quark mass $m_s = 0.37$ GeV, this decay constant would require $\beta_{K^*_0} > 0.60$,
which is quite high for $p$-wave mesons. Particularly, it would also spoil
the matching for decay constants of axial vector $K_1$ mesons which seem to
require $\beta < 0.4 $. Furthermore, choosing $\beta_{K^*_0} > 0.60$, would also enhance
$f_{D^*_{s0}}$ and $f_{D^{1/2}_{s1}}$ unduly high if their $\beta$'s
are taken to be greater than or equal to that of ${K^*_0}$.
The new choice of $m_s = 0.45$ GeV has obviously resulted in
difference in the obtained form factors involving strange mesons
from that given in the earlier work ~\cite{Cheng04}.
(xviii) In this work, we have investigated the variation of the form factors
and their slope parameters for $q^2$ dependence with the variation of
$\beta$ values. Wherever the experimental information is available, we have
used that to fix the errors in the beta values,
otherwise a standard $10 \%$ uncertainty in $\beta$ is assigned to
the remaining $s$-wave and $p$-wave mesons.
\section{Covariant model analysis of form factors}
In this section we first describe the form factors
for $s$-wave mesons within the framework of the covariant
light-front quark model \cite{Jaus99} and then extend it to the
$p$-wave meson case followed by numerical results and discussions in the next section.
\subsection{Form factors for $s$-wave to $s$-wave transitions}
Form factors for $P\to P,V$ transitions are defined by
\begin{eqnarray}
\langle P(P^{\prime\prime})|V_\mu|P(P^\prime)\rangle
&=& P_\mu f_+(q^2) + q_\mu f_-(q^2),
\nonumber\\
\langle V(P^{\prime\prime},\varepsilon^{\prime\prime})|V_\mu|P(P^\prime)\rangle
&=&\epsilon_{\mu\nu\alpha \beta}\,\varepsilon^{{\prime\prime}*\nu}P^\alpha q^\beta\, g({q^2}),
\nonumber\\
\langle V(P^{\prime\prime},\varepsilon^{\prime\prime})|A_\mu|P(P^\prime)\rangle
&=&-i\left\{\varepsilon_\mu^{{\prime\prime}*} f({q^2})
+\varepsilon^{*{\prime\prime}}\cdot P \left[P_\mu a_+({q^2})+q_\mu a_-({q^2})\right]\right\},
\label{eq:ffs}
\end{eqnarray}
where $P=P^\prime+P^{\prime\prime}$, $q=P^\prime-P^{\prime\prime}$ and the convention
$\epsilon_{0123}=1$ is adopted. These form factors are related to
the commonly used Bauer-Stech-Wirbel (BSW) form factors \cite{BSW}
via
\begin{eqnarray} \label{eq:ffsdimless}
F_1^{PP}(q^2)&=&f_+(q^2),\quad
F_0^{PP}(q^2)=f_+(q^2)+\frac{q^2}{q\cdot P} f_-(q^2),
\nonumber\\
V^{PV}(q^2)&=&-(M^\prime+M^{\prime\prime})\, g(q^2),\quad
A_1^{PV}(q^2)=-\frac{f(q^2)}{M^\prime+M^{\prime\prime}},
\nonumber\\
A_2^{PV}(q^2)&=&(M^\prime+M^{\prime\prime})\, a_+(q^2),\quad
A_3^{PV}(q^2)-A_0^{PV}(q^2)=\frac{q^2}{2 M^{\prime\prime}}\, a_-(q^2),
\end{eqnarray}
where the latter form factors are defined by \cite{BSW}
\begin{eqnarray} \label{eq:ffPPV}
\langle P(P^{\prime\prime})|V_\mu|P(P^\prime)\rangle &=& \left(P_\mu-{M^{\prime 2}-M^{{\prime\prime} 2}\over q^2}\,q_ \mu\right)
F_1^{PP}(q^2)+{M^{\prime 2}-M^{{\prime\prime} 2}\over q^2}\,q_\mu\,F_0^{PP}(q^2), \nonumber \\
\langle V(P^{\prime\prime},\varepsilon^{\prime\prime})|V_\mu|P(P^\prime)\rangle &=& -{1\over
M'+M^{\prime\prime}}\,\epsilon_{\mu\nu\alpha \beta}\varepsilon^{{\prime\prime}*\nu}P^\alpha
q^\beta V^{PV}(q^2), \nonumber \\
\langle V(P^{\prime\prime},\varepsilon^{\prime\prime})|A_\mu|P(P^\prime)\rangle &=& i\Big\{
(M'+M^{\prime\prime})\varepsilon^{{\prime\prime}*}_\mu A_1^{PV}(q^2)-{\varepsilon^{{\prime\prime}*}\cdot P\over
M'+M^{\prime\prime}}\,P_\mu A_2^{PV}(q^2) \nonumber \\
&& -2M^{\prime\prime}\,{\varepsilon^{{\prime\prime}*}\cdot P\over
q^2}\,q_\mu\big[A_3^{PV}(q^2)-A_0^{PV}(q^2)\big]\Big\},
\end{eqnarray}
with $F_1^{PP}(0)=F_0^{PP}(0)$, $A_3^{PV}(0)=A_0^{PV}(0)$, and
\begin{eqnarray}
A_3^{PV}(q^2)=\,{M'+M^{\prime\prime}\over
2M^{\prime\prime}}\,A_1^{PV}(q^2)-{M'-M^{\prime\prime}\over 2M^{\prime\prime}}\,A_2^{PV}(q^2).
\end{eqnarray}
Besides the dimensionless form factors,
this parametrization has the advantage that the $q^2$ dependence
of the form factors is governed by the resonances of the same
spin, for instance, the momentum dependence of $F_0(q^2)$ is
determined by scalar resonances.
To obtain the $P\to M$ transition form factors with $M$ being a
ground-state $s$-wave meson (or a low-lying $p$-wave meson), we
follow \cite{Jaus99, Cheng04} to obtain $P\to P,V$ form factors before
considering the $p$-wave meson case.
For the case of $M=P$, it is straightforward to obtain the form
factors $f_\pm(q^2)$ for $q^2=-q_\bot^2\leq0$. We will return to
the issue of the momentum dependence of form factors in the last sub-section. At $q^2=0$,
the form factor $f_+(0)$ is simply given by
\begin{eqnarray} \label{eq:fplus}
f_+(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot
\frac{h^\prime_P h^{\prime\prime}_P}{x_2 \hat N_1^\prime \hat N^{\prime\prime}_1}
\Bigg[x_1 (M_0^{\prime2}+M_0^{\pp2})+x_2 q^2
\nonumber\\
&&\qquad-x_2(m_1^\prime-m_1^{\prime\prime})^2 -x_1(m_1^\prime-m_2)^2-x_1(m_1^{\prime\prime}-m_2)^2\Bigg].
\nonumber\\
\end{eqnarray}
Similarly, we have
\begin{eqnarray}
f_-(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot
\frac{2h^\prime_P h^{\prime\prime}_P}{x_2 \hat N_1^\prime \hat N^{\prime\prime}_1}
\Bigg\{- x_1 x_2 M^{\prime2}-p_\bot^{\prime2}-m_1^\prime m_2
+(m_1^{\prime\prime}-m_2)(x_2 m_1^\prime+x_1 m_2)
\nonumber\\
&&\qquad +2\frac{q\cdot P}{q^2}\left(p^{\prime2}_\bot+2\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}\right)
+2\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}
-\frac{p^\prime_\bot\cdot q_\bot}{q^2}
\Big[M^{\pp2}-x_2(q^2+q\cdot P)
\nonumber\\
&&\qquad -(x_2-x_1) M^{\prime2}+2 x_1 M_0^{\prime
2}-2(m_1^\prime-m_2)(m_1^\prime+m_1^{\prime\prime})\Big]
\Bigg\}.
\label{eq:fpm}
\end{eqnarray}
We next turn to the $P\to V$ transition form factors, which are given by
\begin{eqnarray} \label{eq:PtoV}
g(q^2)&=&-\frac{N_c}{16\pi^3}\int dx_2 d^2 p^\prime_\bot
\frac{2 h^\prime_P h^{\prime\prime}_V}{x_2 \hat N^\prime_1 \hat N^{\prime\prime}_1}
\Bigg\{x_2 m_1^\prime+x_1 m_2+(m_1^\prime-m_1^{\prime\prime})
\frac{p^\prime_\bot\cdot q_\bot}{q^2}
+\frac{2}{w^{\prime\prime}_V}\left[p^{\prime2}_\bot+\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}\right]
\Bigg\},
\nonumber\\
f(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2 p^\prime_\bot
\frac{ h^\prime_P h^{\prime\prime}_V}{x_2 \hat N^\prime_1 \hat N^{\prime\prime}_1}
\Bigg\{2
x_1(m_2-m_1^\prime)(M^{\prime2}_0+M^{\pp2}_0)-4 x_1
m_1^{\prime\prime} M^{\prime2}_0+2x_2 m_1^\prime q\cdot P
\nonumber\\
&&+2 m_2 q^2-2 x_1 m_2
(M^{\prime2}+M^{\pp2})+2(m_1^\prime-m_2)(m_1^\prime+m_1^{\prime\prime})^2
+8(m_1^\prime-m_2)\left[p^{\prime2}_\bot+\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}\right]
\nonumber\\
&&
+2(m_1^\prime+m_1^{\prime\prime})(q^2+q\cdot
P)\frac{p^\prime_\bot\cdot q_\bot}{q^2}
-4\frac{q^2 p^{\prime2}_\bot+(p^\prime_\bot\cdot q_\bot)^2}{q^2 w^{\prime\prime}_V}
\Bigg[2 x_1 (M^{\prime2}+M^{\prime2}_0)-q^2-q\cdot P
\nonumber\\
&&-2(q^2+q\cdot P)\frac{p^\prime_\bot\cdot
q_\bot}{q^2}-2(m_1^\prime-m_1^{\prime\prime})(m_1^\prime-m_2)
\Bigg]\Bigg\},
\nonumber\\
a_+(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2 p^\prime_\bot
\frac{2 h^\prime_P h^{\prime\prime}_V}{x_2 \hat N^\prime_1 \hat N^{\prime\prime}_1}
\Bigg\{(x_1-x_2)(x_2 m_1^\prime+x_1 m_2)-[2x_1
m_2+m_1^{\prime\prime}+(x_2-x_1)
m_1^\prime]\frac{p^\prime_\bot\cdot q_\bot}{q^2}
\nonumber\\
&&-2\frac{x_2 q^2+p_\bot^\prime\cdot q_\bot}{x_2 q^2
w^{\prime\prime}_V}\Big[p^\prime_\bot\cdot p^{\prime\prime}_\bot+(x_1 m_2+x_2 m_1^\prime)(x_1 m_2-x_2 m_1^{\prime\prime})\Big]
\Bigg\},
\nonumber\\
a_-(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2 p^\prime_\bot
\frac{ h^\prime_P h^{\prime\prime}_V}{x_2 \hat N^\prime_1 \hat N^{\prime\prime}_1}
\Bigg\{2(2x_1-3)(x_2 m_1^\prime+x_1
m_2)-8(m_1^\prime-m_2)\left[\frac{p^{\prime2}_\bot}{q^2}+2\frac{(p^\prime_\bot\cdot
q_\bot)^2}{q^4}\right]
\nonumber\\
&&-[(14-12 x_1) m_1^\prime-2 m_1^{\prime\prime}-(8-12 x_1) m_2]
\frac{p^\prime_\bot\cdot q_\bot}{q^2}
\nonumber\\
&&+\frac{4}{w^{\prime\prime}_V}\Bigg([M^{\prime2}+M^{\pp2}-q^2+2(m_1^\prime-m_2)(m_1^{\prime\prime}+m_2)]
(A^{(2)}_3+A^{(2)}_4-A^{(1)}_2)
\nonumber\\
&& +Z_2(3 A^{(1)}_2-2A^{(2)}_4-1)
+\frac{1}{2}[x_1(q^2+q\cdot P)
-2 M^{\prime2}-2 p^\prime_\bot\cdot q_\bot
\nonumber\\
&& -2 m_1^\prime(m_1^{\prime\prime}+m_2)-2
m_2(m_1^\prime-m_2)
](A^{(1)}_1+A^{(1)}_2-1)
\nonumber\\
&& +q\cdot P\Bigg[\frac{p^{\prime2}_\bot}{q^2}
+\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^4}\Bigg]
(4
A^{(1)}_2-3)
\Bigg)
\Bigg\},
\end{eqnarray}
where various quantities appearing in these formulas have been described in the previous section.
\subsection{Form factors for $s$-wave to $p$-wave transitions}
The general expressions for $P$ to low-lying $p$-wave meson
transitions are given by \cite{ISGW}
\begin{eqnarray} \label{ISGWform}
\langle S(P^{\prime\prime})|A_\mu|P(P^\prime)\rangle &=& i\Big[ u_+(q^2)P_\mu+u_-(q^2)q_\mu
\Big], \nonumber \\
\langle A^{1/ 2}(P^{\prime\prime},\varepsilon^{\prime\prime})|V_\mu|P(P^\prime)\rangle
&=& i\left\{\ell_{1/2}(q^2)\varepsilon_\mu^{{\prime\prime}*}+\varepsilon^{{\prime\prime}*}\cdot
P[P_\mu c_+^{1/2}(q^2)+q_\mu c_-^{1/2}(q^2)]\right\},
\nonumber \\
\langle A^{1/2}(P^{\prime\prime},\varepsilon^{\prime\prime})|A_\mu|P(P^\prime)\rangle
&=& -q_{1/2}(q^2)\epsilon_{\mu\nu\alpha\beta}\varepsilon^{{\prime\prime}*\nu}P^\alpha
q^\beta,
\nonumber \\
\langle A^{3/ 2}(P^{\prime\prime},\varepsilon^{\prime\prime})|V_\mu|P(P^\prime)\rangle
&=& i\left\{\ell_{3/2}(q^2)\varepsilon_\mu^{{\prime\prime}*}+\varepsilon^{{\prime\prime}*}\cdot
P[P_\mu c_+^{3/2}(q^2)+q_\mu c_-^{3/2}(q^2)]\right\},
\nonumber \\
\langle A^{3/2}(P^{\prime\prime},\varepsilon^{\prime\prime})|A_\mu|P(P^\prime)\rangle
&=& -q_{3/2}(q^2)\epsilon_{\mu\nu\alpha\beta}\varepsilon^{{\prime\prime}*\nu}P^\alpha
q^\beta.
\label{eq:ffp}
\end{eqnarray}
The form factors $\ell_{1/2(3/2)},c_+^{1/2(3/2)},c_-^{1/2(3/2)}$
and $q_{1/2(3/2)}$ are defined for the transitions to the heavy
$P^{1/2}_1$ ($P^{3/2}_1$) state. For transitions to light
axial-vector mesons, it is more appropriate to employ the $L-S$
coupled states $^1P_1$ and $^3P_1$ denoted by the particles
$^1\!A$ and $^3\!A$ in our notation. The relation between
$P^{1/2}_1,P^{3/2}_1$ and $^1P_1,\,^3P_1$ states is given by Eq.
(\ref{eq:Phalf}). The corresponding form factors
$\ell_{^1\!A(^3\!A)},c_+^{^1\!A(^3\!A)},c_-^{^1\!A(^3\!A)}$ and
$q_{^1\!A(^3\!A)}$ for $P\to \,^1\!A$ ($^3\!A$) transitions can be
defined in an analogous way.\footnote{The form factors
$\ell_{^1\!A(^3\!A)},c_+^{^1\!A(^3\!A)},c_-^{^1\!A(^3\!A)}$ and
$q_{^1\!A(^3\!A)}$ are dubbed as $\ell(v),c_+(s_+),c_-(s_-)$ and
$q(r)$, respectively, in the ISGW model \cite{ISGW}.}
Note that only the form factors $u_+(q^2),u_-(q^2)$ and $k(q^2)$
in the above parametrization are dimensionless. It is thus
convenient to define dimensionless form factors by\footnote{The
definition here for dimensionless $P\to A$ transition form factors
differs than Eq. (3.17) of \cite{Cheng:2003id} where the
coefficients $(m_P\pm m_A)$ are replaced by $(m_P\mp m_A)$. It
has been made clear in the earlier work \cite{Cheng04}
that this definition will lead to HQS
relations for $B\to D_0^*,D_1$ transitions similar to that for $B\to D,D^*$ ones.}
\begin{eqnarray} \label{eq:ffpdimless}
\langle S(P^{\prime\prime})|A_\mu|P(P^\prime)\rangle &=&
-i\left[\left(P_\mu-{M^{\prime 2}-M^{{\prime\prime} 2}\over q^2}\,q_
\mu\right) F_1^{PS}(q^2)+{M^{\prime 2}-M^{{\prime\prime} 2}\over q^2}
\,q_\mu\,F_0^{PS}(q^2)\right], \nonumber \\
\langle A(P^{\prime\prime},\varepsilon^{\prime\prime})|V_\mu|P(P^\prime)\rangle &=&
-i\Bigg\{(m_P-m_A) \varepsilon^*_\mu V_1^{PA}(q^2) - {\varepsilon^*\cdot
P'\over m_P-m_A}P_\mu V_2^{PA}(q^2) \nonumber \\
&-& 2m_A {\varepsilon^*\cdot P'\over
q^2}q_\mu\left[V_3^{PA}(q^2)-V_0^{PA}(q^2)\right]\Bigg\},
\nonumber \\
\langle A(P^{\prime\prime},\varepsilon^{\prime\prime})|A_\mu|P(P^\prime)\rangle &=& -{1\over
m_P-m_A}\,\epsilon_{\mu\nu\rho\sigma}\varepsilon^{*\nu}P^\rho
q^{\sigma}A^{PA}(q^2),
\end{eqnarray}
with
\begin{eqnarray} V_3^{PA}(q^2)=\,{m_P-m_A\over 2m_A}\,V_1^{PA}(q^2)-{m_P+m_A\over
2m_A}\,V_2^{PA}(q^2),
\end{eqnarray}
and $V_3^{PA}(0)=V_0^{PA}(0)$. They are related to the form
factors in (\ref{eq:ffPPV}) via
\begin{eqnarray}
F^{PS}_1(q^2)&=&-u_+(q^2),\quad
F^{PS}_0(q^2)=-u_+(q^2)-\frac{q^2}{q\cdot P} u_-(q^2),
\nonumber\\
A^{PA}(q^2)&=&-(M^\prime-M^{\prime\prime})\, q(q^2),\quad
V^{PA}_1(q^2)=-\frac{\ell(q^2)}{M^\prime-M^{\prime\prime}},
\nonumber\\
V^{PA}_2(q^2)&=&(M^\prime-M^{\prime\prime})\, c_+(q^2),\quad
V^{PA}_3(q^2)-V^{PA}_0(q^2)=\frac{q^2}{2 M^{\prime\prime}}\, c_-(q^2).
\end{eqnarray}
In above equations, the axial-vector meson $A$ stands for
$A^{1/2}$ or $A^{3/2}$.
The $P\to S\,(A)$ transition form factors can be easily
obtained by some suitable modifications on $P\to P\,(V)$ ones.
The $P\to S$ transition form factors are related to $f_\pm$ by
\begin{eqnarray}
u_\pm=-f_\pm(m_1^{\prime\prime}\to -m_1^{\prime\prime},h^{\prime\prime}_P\to h^{\prime\prime}_S).
\end{eqnarray}
Thus the following form of these form factors can be obtained
from that of $P\to P$ ones by the replacements given above,
\begin{eqnarray}
u_+(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot
\frac{h^\prime_P h^{\prime\prime}_S}{x_2 \hat N_1^\prime \hat N^{\prime\prime}_1}
\Big[-x_1 (M_0^{\prime2}+M_0^{\pp2})-x_2 q^2
\nonumber\\
&&\qquad+x_2(m_1^\prime+m_1^{\prime\prime})^2 +x_1(m_1^\prime-m_2)^2+x_1(m_1^{\prime\prime}+m_2)^2\Big],
\nonumber\\
u_-(q^2)&=&\frac{N_c}{16\pi^3}\int dx_2 d^2p^\prime_\bot
\frac{2h^\prime_P h^{\prime\prime}_S}{x_2 \hat N_1^\prime \hat N^{\prime\prime}_1}
\Bigg\{ x_1 x_2 M^{\prime2}+p_\bot^{\prime2}+m_1^\prime m_2
+(m_1^{\prime\prime}+m_2)(x_2 m_1^\prime+x_1 m_2)
\nonumber\\
&&\qquad -2\frac{q\cdot P}{q^2}\left(p^{\prime2}_\bot+2\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}\right)
-2\frac{(p^\prime_\bot\cdot q_\bot)^2}{q^2}
+\frac{p^\prime_\bot\cdot q_\bot}{q^2}
\Big[M^{\pp2}-x_2(q^2+q\cdot P)
\nonumber\\
&&\qquad -(x_2-x_1) M^{\prime2}+2 x_1 M_0^{\prime
2}-2(m_1^\prime-m_2)(m_1^\prime-m_1^{\prime\prime})\Big]
\Bigg\}.
\label{eq:upm}
\end{eqnarray}
Similarly, the analytic expressions for $P\to A$ transition form factors
can be obtained from that of $P\to V$ ones by the following
replacements:
\begin{eqnarray} \label{eq:PtoA}
\ell^{^3\!A,^1\!A}(q^2)&=&f(q^2) \,\,\,{\rm with}\,\,\,
(m_1^{\prime\prime}\to -m_1^{\prime\prime},\,h^{\prime\prime}_V\to h^{\prime\prime}_{^3\!A,^1\!A},\,w^{\prime\prime}_V\to w^{\prime\prime}_{^3\!A,^1\!A}),
\nonumber\\
q^{^3\!A,^1\!A}(q^2)&=&g(q^2) \,\,\,{\rm with}\,\,\,
(m_1^{\prime\prime}\to -m_1^{\prime\prime},\,h^{\prime\prime}_V\to h^{\prime\prime}_{^3\!A,^1\!A},\,w^{\prime\prime}_V\to w^{\prime\prime}_{^3\!A,^1\!A}),
\nonumber\\
c_+^{^3\!A,^1\!A}(q^2)&=&a_+(q^2) \,\,\,{\rm with}\,\,\,
(m_1^{\prime\prime}\to -m_1^{\prime\prime},\,h^{\prime\prime}_V\to h^{\prime\prime}_{^3\!A,^1\!A},\,w^{\prime\prime}_V\to w^{\prime\prime}_{^3\!A,^1\!A}),
\nonumber\\
c_-^{^3\!A,^1\!A}(q^2)&=&a_-(q^2) \,\,\,{\rm with}\,\,\,
(m_1^{\prime\prime}\to -m_1^{\prime\prime},\,h^{\prime\prime}_V\to h^{\prime\prime}_{^3\!A,^1\!A},\,w^{\prime\prime}_V\to w^{\prime\prime}_{^3\!A,^1\!A}).
\end{eqnarray}
It should be cautious that the replacement of $m_1^{\prime\prime}\to
-m_1^{\prime\prime}$ should not be applied to $m_1^{\prime\prime}$ in $w^{\prime\prime}$ and
$h^{\prime\prime}$. These form factors can be expressed in the $P^{3/2}_1$
and $P^{1/2}_1$ basis by using Eq.~(\ref{eq:Phalf}). For further details,
the reader is referred to the earlier work \cite{Cheng04}.
\subsection{Form-factor momentum dependence and numerical results}
Because of the condition $q^+=0$ we have imposed during the course
of calculation, form factors are known only for spacelike momentum
transfer $q^2=-q^2_\bot\leq 0$, whereas only the timelike form
factors are relevant for the physical decay processes. It has been
proposed in \cite{Jaus96} to recast the form factors as explicit
functions of $q^2$ in the spacelike region and then analytically
continue them to the timelike region. Another approach is to
construct a double spectral representation for form factors at
$q^2<0$ and then analytically continue it to $q^2>0$
region~\cite{Melikhov96}. It has been shown recently that, within
a specific model, form factors obtained directly from the timelike
region (with $q^+>0$) are identical to the ones obtained by the
analytic continuation from the spacelike region~\cite{BCJ03}.
In principle, form factors at $q^2>0$ can be evaluated directly in
the frame where the momentum transfer is purely longitudinal,
i.e., $q_\bot=0$, so that $q^2=q^+q^-$ covers the entire range of
momentum transfer \cite{Cheng97}. The price one has to pay is
that, besides the conventional valence-quark contribution, one
must also consider the non-valence configuration (or the so-called
$Z$-graph) arising from quark-pair creation from the vacuum.
However, a reliable way of estimating the $Z$-graph contribution
is still lacking unless one works in a specific model, for
example, the one advocated in \cite{BCJ03}. Fortunately, this
additional non-valence contribution vanishes in the frame where
the momentum transfer is purely transverse i.e., $q^+=0$.
To proceed we find that, except for the form factor $V_2$ to be
discussed below, the momentum dependence of form factors in the
spacelike region can be well parameterized and reproduced in the following
three-parameter form:
\begin{eqnarray}
\label{eq:FFpara}
F(q^2)= { F(0)\over {1-a(q^2/m_{B(D)}^2)+b(q^2/m_{B(D)}^2)^2}} ,
\end{eqnarray}
for $P \to M$ transitions, where $F$ stands for the relevant form factors
appearing in these transitions.
The parameters $a$, $b$ and $F(0)$ are first determined in the
spacelike region. We then employ this parametrization to determine
the physical form factors at $q^2\geq 0$. In practice, the
parameters $a,b$ and $F(0)$ are obtained by performing a
5-parameter fit to the form factors in the range $-20\,{\rm
GeV}^2\leq q^2\leq 0$ for $B$ decays and $-10\,{\rm GeV}^2\leq
q^2\leq 0$ for $D$ decays. All $P \rightarrow M$ form factors are
calculated at five $q^2$ values given below:
a) for the charm sector: $q^2 = -0.01, -0.1, -1.0, -5.0, -10.0 ~ \textrm{GeV}^2 $,
b) for the bottom sector: $q^2 = -0.01, -0.1, -5.0, -10.0, -20.0 ~ \textrm{GeV}^2 $.
These parameters are generally
insensitive to the $q^2$ range to be fitted except for the form
factor $V_2(q^2)$ in $B(D)\to\, ^1P_1,P_1^{3/2}$ transitions.
The obtained $a$ and $b$ coefficients are in most
cases not far from unity as expected.
We have also analyzed the sensitivity of the form factors $F(0)$, and the
slope parameters ($a$ and $b$) to the uncertainties
of $\beta$ values. The form factors at $q^2 = 0 $ are generally found to be
less sensitive to
the variation in $\beta$ values, whereas the corresponding parameters $a$ and $b$
are rather sensitive to the chosen range for $\beta$. Numerical results
and discussion of these form factors and slope parameters ($a$ and $b$)
are presented in detail in the following section.
\section{Numerical results and discussion}
Equipped with the explicit expressions of the form
factors $f_+(q^2),f_-(q^2)$ [ Eqs. (\ref{eq:fplus}) and (\ref{eq:fpm})] for $P\to P$
transitions, $g(q^2),f(q^2),a_+(q^2),a_-(q^2)$ [Eq.
(\ref{eq:PtoV})] for $P\to V$ transitions,
$u_+(q^2),u_-(q^2)$ [Eq. (\ref{eq:upm})] for $P\to S$
transitions, and
$\ell(q^2),q(q^2),c_+(q^2),c_-(q^2)$ [Eq. (\ref{eq:PtoA})] for
$P\to A$ transition, we now proceed to perform their
numerical studies. In the earlier work, results for the
form factors for $D (c \bar u)$ ,
$B(b \bar u) \to $ isovector ($\pi$ like) and isospinor
($K$ and $D$ like) transition were calculated. In this work,
we include isoscalar initial and final state mesons as well. Besides giving the updated
results of these transitions due to the change in the strange
quark mass, the variation of the $\beta$ values and performing fit for five $q^2$
values, the $D_s, B_s \to P, V, S$, and $A$ transition form factors are the
main new results in this work.
In Tables~\ref{tab:P2P}$-$\ref{tab:B2D},
we present calculated form factors and their $q^2$ dependence, along with their
allowed range due to uncertainties in $\beta$ values of the initial
and final mesons, for the $P (0^-) \to P(0^-), V(1^-), S(0+), A(1^+ ~ ^3P_1)$,
and $A(1^+ ~ ^1P_1)$ transitions of the charmed $D, D_s$ and bottom $B, B_s$ mesons.
In calculations, we have taken the meson masses from the Particle Data Group
\cite{PDG10}. Taking the natural flavor basis for isoscalar states of all the mesons (M), i.e.,
$M_{q} = (u\overline{u} + d\overline{d})/\surd {2}$ and $M_s = (s\overline{s})$,
we use the following masses (in GeV): $m_{\eta_q} = 0.741$ and $m_{\eta_s} = 0.802$
for pseudoscalar mesons taken from an analysis given in \cite{ChengChua10},
$m_{f_{1q}} = 1.283$ and $m_{f_{1s}} = 1.425$ for the axial-vector ($1^{++}$) nonet,
$m_{h_{1q}} = 1.242$ and $m_{h_{1s}} = 1.314$ for the other axial-vector ($1^{+-}$) case, and
$m_{f_{0q}} = 1.474$ and $m_{f_{0s}} = 1.5$
for the scalar ($0^{++}$) mesons,
based on the respective mixing schemes described in Sec. III. Form factors for
transitions to the physical isosinglet diagonal states
can be obtained from the Tables by including suitable Clebsch-Gordan coefficients.
For instance,
\begin{eqnarray} \label{eq:FFf1CG}
A^{B_s f_1 (1420)} = - \sin \alpha_{f_1} A^{B_s f_{1s}},~~
\nonumber
A^{B_s f_1 (1285)} = \cos\alpha_{f_1} A^{B_s f_{1s}},
\\
A^{B f_1 (1420)} = {1 \over {\sqrt{2}}} \cos \alpha_{f_1} A^{B f_{1q}},~~
A^{B f_1 (1285)} = {1 \over {\sqrt{2}}} \sin \alpha_{f_1} A^{B f_{1q}},
\end{eqnarray}
where $\alpha_{f_1}$ has already been defined in Sec. III. The factor $\sqrt{2}$ appears
for the $B\to f_1$ form factors, since either $u \overline{u}$ or
$d \overline{d}$ component of $f_{1q}$
can be transited from $B$ meson via the appropriate weak current. Similarly,
only the $s \overline{s}$ components of these mesons can be transited from the $B_s$
meson. So the size of these corresponding form factors for
physical isoscalar diagonal states gets reduced by the Clebsch-Gordan coefficients. Similar procedure
can be adopted for transitions to the isosinglet diagonal states in other multiplets.
In these tables, two sets of uncertainties in the form factors,
commonly denotes as $F(0)$, and their slope parameters ($a$ and $b$) are
given. The first and second sets of uncertainties
shown in their values arise from the allowed uncertainties in the $\beta $ parameter of
the initial and final state meson, respectively. For the sake of clarity, it is mentioned
here that the uncertainty shown as superscript (subscript) is due to the
increase (decrease) in $\beta $ of the corresponding meson.
The obtained $a$ and $b$ coefficients are in most
cases not far from unity as expected. These parameters are generally
insensitive to the $q^2$ range to be fitted, except for the form factor
$V_2(q^2)$ in $B(D)\to\, ^1P_1,P_1^{3/2}$ transitions. For these transitions, the
corresponding parameters $a$ and $b$ are rather sensitive to the
chosen range for $q^2$, and quite larger than unity. This sensitivity is attributed to the fact
that the form factor $V_2(q^2)$ approaches to zero at very large
$-|q^2|$ where the three-parameter parametrization
(\ref{eq:FFpara}) becomes questionable. To overcome this
difficulty, we follow \cite{Cheng04} to fit this form factor to the following form:
\begin{eqnarray} \label{eq:FFpara1}
F(q^2)=\,{F(0)\over (1-q^2/m_{B(D)}^2)[1-a(q^2/m_{B(D)}^2)+b(q^2/m_{B(D)}^2)^2]},
\end{eqnarray}
and achieve a substantial improvement. For example, we have
$a=2.18$ and $b=6.08$ when $V_2^{BK_{^1P_1}}$ is fitted to Eq.
(\ref{eq:FFpara}) and they become $a=1.74$ and $b=2.17$ (see Table
\ref{tab:P2An}) when the fit formula Eq. (\ref{eq:FFpara1}) is
employed. It may be noted that we have considered parent meson constituted of
heavy quark and light antiquark, since certain decay constants and form factors
may change sign.
We make the following observations:
\subsection{ $P(0^-) \to P (0^-)$ Form Factors}
\begin{itemize}
\item From Table~\ref{tab:P2P}, we notice that heavy-to-light form factors
for the bottom mesons are smaller (around 0.3) than all
the charmed meson form factors and the heavy-to-heavy bottom meson form
factors, $F^{BD(B_sD_s)}_{0,1}$, which are around 0.7 or 0.8.
\item We notice that the values of form factors at $q^2= 0$ for $B_s$ transitions
are similar to the corresponding ones in $B$ transitions. Therefore, flavor of
the spectator quark does not seem to play a special role in affecting them.
Particularly, we note the following for both $F_0(0)$ and $F_1(0)$:
$F^{B_s D_s} = F^{B D}$, $F^{B_s K}\approx F^{B \pi}$, and
$F^{B_s \eta_s}\approx F^{B \eta_q}$, where
$\eta_{q}= (u\overline{u} + d\overline{d})/\surd {2}$, and
$\eta_{s}$ is pure $(s\overline{s})$ state. For the charm sector also, one may
notice $F^{D_s K} = F^{D \pi} (\approx F^{BD})$ and $F^{D_s \eta_s}\approx F^{D K}$.
However, the slope parameters, $a$ and $b$, differ for these cases.
\item Since the decay constants of pseudoscalar mesons are quite accurately
determined, the errors on the $\beta$ parameters are rather small.
Correspondingly, the errors in the calculated form factors at $q^2 = 0$ are
also very small. The same is true for the slope parameters except for a
few cases, particularly for $b$, which may show large variation sometimes.
\item Form factors $(F^{PP}_1(0)$ and $F^{PP}_0(0))$ usually tend to decrease (increase)
with increasing (decreasing) $\beta$ for initial meson, whereas they tend to increase (decrease)
with increasing (decreasing) $\beta$ for final meson. Only for $B_s$, the form factors
show increasing (decreasing) trend for the initial as well as the final meson.
\item Usually all the slope parameters are found to be positive.
For the bottom sector, the slope parameters are larger than that for the charm sector.
Particularly, the parameter $b$ is much small ($< 0.1$, if not zero) for $F^{PP}_0(0)$, except for
$F^{B_s K}_0(0)$ and $F^{B_s \eta_s}_0(0)$ for which $b \approx 0.35$. For
$F^{B_s K}_1(0)$ and $F^{B_s \eta_s}_1(0)$, the parameter $b$ is around $2-3$
times larger than that for other cases.
\item Slope parameters obtained using Eq. (\ref{eq:FFpara}) generally
tend to increase (decrease) with decrease (increase) in $\beta$ for
each of the initial and final mesons.
\item According to the three-parameter parametrization Eq. (\ref{eq:FFpara}), the dipole behavior
corresponds to $b=(a/2)^2$, while $b=0$ and $a\neq0$ induces
a monopole dependence. An inspection of Table~\ref{tab:P2P}
indicates that form factors $F^{PP}_0$ generally show a monopole behavior, and $F^{PP}_1$
have a dipole behavior particularly for charmed meson transitions.
\end{itemize}
\begin{table}[b]
\caption{Form factors of
$P(0^-)\rightarrow P(0^-)$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless. }
\label{tab:P2P}
\begin{ruledtabular}
\begin{tabular}{|cccc|cccc|}
$F$
& $F(0)$
& $a$
& $b$
& $F$
& $F(0)$
& $a$
& $b$
\\
\hline
$F^{D \pi}_1$
& $0.66^{-0.01+0.00}_{+0.01-0.00}$
& $1.19^{-0.01-0.00}_{+0.01+0.00}$
& $0.35^{-0.03-0.00}_{+0.03+0.00}$
& $F^{D \pi}_0$
& $0.66^{-0.01+0.00}_{+0.01-0.00}$
& $0.51^{-0.00-0.00}_{+0.00+0.00}$
& $0.00^{-0.01-0.00}_{+0.01+0.00}$
\\
$F^{D {\eta}_q}_1$
& $0.71^{-0.00+0.01}_{+0.00-0.01}$
& $1.13^{-0.01-0.02}_{+0.01+0.02}$
& $0.27^{-0.02-0.02}_{+0.02+0.02}$
& $F^{D {\eta}_q}_0$
& $0.71^{-0.01+0.01}_{+0.00-0.01}$
& $0.43^{+0.01-0.03}_{-0.01+0.03}$
& $-0.01^{-0.00+0.00}_{+0.01-0.00}$
\\
$F^{DK}_1$
& $0.79^{-0.01+0.00}_{+0.01-0.00}$
& $1.05^{-0.01-0.00}_{+0.01+0.00}$
& $0.25^{-0.02-0.00}_{+0.02+0.00}$
& $F^{DK}_0$
& $0.79^{-0.01+0.00}_{+0.01-0.00}$
& $0.47^{+0.00-0.01}_{-0.00+0.01}$
& $-0.00^{-0.00-0.00}_{+0.00+0.00}$
\\
\hline
$F^{D_s K}_1$
& $0.66^{-0.00+0.00}_{+0.00-0.00}$
& $1.11^{-0.00-0.00}_{+0.00+0.00}$
& $0.48^{-0.03-0.01}_{+0.03+0.01}$
& $F^{D_s K}_0$
& $0.66^{-0.00+0.00}_{+0.00-0.00}$
& $0.56^{-0.00-0.01}_{+0.00+0.01}$
& $0.04^{-0.01-0.00}_{+0.01+0.00}$
\\
$F^{D_s {\eta}_s}_1$
& $0.76^{-0.00+0.02}_{+0.00-0.03}$
& $1.02^{-0.01-0.01}_{+0.00+0.01}$
& $0.40^{-0.02-0.05}_{+0.02+0.05}$
& $F^{D_s {\eta}_s}_0$
& $0.76^{-0.01+0.02}_{+0.00-0.03}$
& $0.60^{-0.00-0.06}_{-0.00+0.05}$
& $0.04^{-0.00-0.01}_{+0.00+0.01}$
\\
\hline
$F^{B\pi}_1$
& $0.25^{-0.00+0.00}_{+0.00-0.00}$
& $1.70^{-0.03-0.00}_{+0.03+0.00}$
& $0.90^{-0.05-0.00}_{+0.06+0.00}$
& $F^{B\pi}_0$
& $0.25^{-0.00+0.00}_{+0.00-0.00}$
& $0.82^{-0.02-0.00}_{+0.02+0.00}$
& $0.09^{-0.01-0.00}_{+0.02+0.00}$
\\
$F^{B {\eta}_q}_1$
& $0.29^{-0.00+0.01}_{+0.00-0.01}$
& $1.63^{-0.02-0.02}_{+0.02+0.02}$
& $0.74^{-0.04-0.04}_{+0.04+0.04}$
& $F^{B {\eta}_q}_0$
& $0.29^{-0.00+0.01}_{+0.00-0.01}$
& $0.75^{-0.01-0.03}_{+0.01+0.03}$
& $0.04^{-0.01-0.01}_{+0.01+0.01}$
\\
$F^{BK}_1$
& $0.34^{-0.00+0.00}_{+0.00-0.00}$
& $1.60^{-0.02-0.00}_{+0.02+0.00}$
& $0.73^{-0.04-0.01}_{+0.04+0.01}$
& $F^{BK}_0$
& $0.34^{-0.00+0.00}_{+0.00-0.00}$
& $0.78^{-0.02-0.01}_{+0.02+0.01}$
& $0.05^{-0.01-0.00}_{+0.01+0.00}$
\\
$F^{BD}_1$
& $0.67^{-0.00+0.01}_{+0.00-0.01}$
& $1.22^{-0.01-0.02}_{+0.01+0.01}$
& $0.36^{-0.01-0.01}_{+0.01+0.02}$
& $F^{BD}_0$
& $0.67^{-0.00+0.01}_{+0.00-0.01}$
& $0.63^{-0.00-0.02}_{+0.00+0.02}$
& $-0.01^{-0.01-0.00}_{+0.01+0.00}$
\\
\hline
$F^{B_s K}_1$
& $0.23^{+0.00+0.00}_{-0.00-0.00}$
& $1.88^{-0.04-0.01}_{+0.04+0.01}$
& $1.58^{-0.12-0.03}_{+0.14+0.03}$
& $F^{B_s K}_0$
& $0.23^{+0.00+0.00}_{-0.00-0.00}$
& $1.05^{-0.03-0.01}_{+0.04+0.01}$
& $0.35^{-0.05-0.02}_{+0.04+0.00}$
\\
$F^{B_s {\eta}_s}_1$
& $0.28^{+0.00+0.02}_{-0.00-0.02}$
& $1.82^{-0.04-0.05}_{+0.04+0.05}$
& $1.45^{-0.11-0.16}_{+0.13+0.18}$
& $F^{B_s {\eta}_s}_0$
& $0.28^{+0.00+0.02}_{-0.00-0.02}$
& $1.07^{-0.03-0.06}_{+0.03+0.07}$
& $0.32^{-0.04-0.06}_{+0.05+0.07}$
\\
$F^{B_s D_s}_1$
& $0.67^{+0.00+0.01}_{-0.01-0.01}$
& $1.28^{-0.02-0.02}_{+0.02+0.02}$
& $0.52^{-0.03-0.02}_{+0.03+0.02}$
& $F^{B_s D_s}_0$
& $0.67^{+0.00+0.01}_{-0.01-0.01}$
& $0.69^{-0.01-0.02}_{+0.00+0.02}$
& $0.07^{-0.01-0.01}_{+0.01+0.01}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
\subsection{$P(0^-) \to V (1^-)$ Form Factors}
\begin{itemize}
\item Like $P(0^-) \to P (0^-)$ form factors, we note
the following behavior from Table~\ref{tab:P2V}:
$F^{B_s D^*_s}\approx F^{B D^*}$, $F^{B_s \phi}\approx F^{B \rho}\approx F^{B \omega}$,
$F^{D_s \phi}\approx F^{D K^*}$ and $F^{D_s K^*}\approx F^{D \rho}\approx F^{D \omega}$,
where $F$ represents any of the form factors, $V^{PV}, A^{PV}_0, A^{PV}_1$ or $A^{PV}_2$.
However, the slope parameters show considerable differences for these cases.
\item It is observed that heavy-to-light
form factors for the bottom mesons are smaller (between 0.2 to 0.4) than all
the charmed meson form factors and the heavy-to-heavy bottom meson form
factors, $F^{BD^*(B_sD^*_s)}$, which lie between 0.6 to 1. We also
notice the pattern $A^{PV}_0 > A^{PV}_1 > A^{PV}_2$ for all transitions,
whereas $V^{PV} > A^{PV}_0$ for charmed meson and $B/B_s \to D/D_s$ transitions,
but $V^{PV}$ is slightly smaller than $A^{PV}_0$ and remains
greater than $A^{PV}_{1,2}$ for the heavy-to-light bottom transitions.
\item Here also due to the reliability in fixing the $\beta$ parameter for lighter vector mesons
and the parent pseudoscalar mesons, the errors in the calculated form factors
$F(0)$ are very small. In contrast, the slope parameters do show
sensitivity to the variation in the $\beta$ parameters specially in the bottom sector.
\item Form factors, $V^{PV}(0), A^{PV}_0(0)$, and $A^{PV}_1(0)$, for the
charm sector usually tend to decrease (increase) with increasing (decreasing) $\beta$
for initial meson, whereas they tend to increase (decrease) with increasing
(decreasing) $\beta$ for final meson. However, the form factor $A^{PV}_2(0)$
shows the opposite trend.
\item For the bottom sector, form factors generally tend to increase (decrease)
with increasing (decreasing) $\beta$ for each of the initial and the final mesons.
\item Slope parameters for all the cases are found to carry positive values. Both parameters
($a$ and $b$) generally
tend to increase (decrease) with decrease (increase) in $\beta$ for
each of the initial and final mesons.
\item The parameters $a$ is usually less sensitive to the $\beta$ variation,
whereas $b$ is more sensitive to $\beta$ values and may show large
variation $(10 \%)$ or even more sometimes for the bottom sector.
\item Almost all the form factors for $D$ as well as $B$ are higher by $(5-10)\%$
than that obtained in the earlier work \cite{Cheng04}, whereas
both slope parameter are reduced in magnitude. This could happen because now we
perform 5-point fit for $q^2$ values.
\item On comparison with $ P \to P, V$ form factors obtained in the BSW model \cite{BSW}, the
Melikhov-Stech (MS) model \cite{Melikhov}, QCD sum rule (QSR) \cite{Ball91},
light-cone sum rules (LCSR) \cite{LCSR} and lattice calculations \cite{LatticeFF},
it is pointed out that our predictions agree well with the available lattice
results, and are most close to that of the MS model except for $B_s$ transitions,
which larger than our results. The LCSR and BSW model results are usually larger
for $P \to V$ form factors for $D$ and $B$ transitions, however
LCSR form factors for $B_s \to K^*$ transition match well with present work. The QSR calculations
are generally lower than our results, except for $B \to K^*$ form factors which are
higher than our predictions. Recently, $P \to V $
form factors for bottom mesons have also been calculated
in the perturbative QCD approach \cite{RHLi09},
which are found to be lower than the values obtained in the present work.
\item Experimentally, the form factors ratios $r_V \equiv V^{PV}(0)/A^{PV}_1(0)$ and
$r_2 \equiv A^{PV}_2(0)/A^{PV}_1(0)$ are available for two semileptonic decays
$D \to K^* \ell \nu$ and $D \to \phi \ell \nu$ \cite{PDG10}:
\begin{eqnarray} \label{eq:DKstar}
r_V(D \to K^*) = 1.62 \pm 0.08, ~~ r_2(D \to K^*) = 0.83 \pm 0.05,
\nonumber\\
r_V(D_s \to \phi) = 1.82 \pm 0.08, ~~ r_2(D_s \to \phi) = 0.84 \pm 0.11.
\end{eqnarray}
Our predictions $r_2(D \to K^*) = 0.83$ and $r_2(D_s \to \phi)= 0.86$,
agree well for both the decays, whereas $r_V(D \to K^*) = 1.36$
and $r_V(D_s \to \phi)= 1.42$ are lower than the corresponding experimental values.
\end{itemize}
\begin{table}[b]
\caption{Form factors of
$P(0^-)\rightarrow V(1^-)$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless. }
\label{tab:P2V}
\begin{ruledtabular}
\begin{tabular}{| c c c c | c c c c |}
$F$
& $F(0)$
& $a$
& $b$
& $F$
& $F(0)$
& $a$
& $b$
\\
\hline
$V^{D \rho}$
& $0.88^{-0.02+0.01}_{+0.01-0.01}$
& $1.23^{-0.01-0.00}_{+0.01+0.00}$
& $0.40^{-0.03-0.01}_{+0.04+0.01}$
& $A^{D \rho}_0$
& $0.69^{-0.01+0.01}_{+0.01-0.01}$
& $1.08^{-0.02-0.00}_{+0.01+0.00}$
& $0.45^{-0.03-0.01}_{+0.03+0.01}$
\\
$A^{D \rho}_1$
& $0.60^{-0.00+0.00}_{+0.00-0.01}$
& $0.46^{-0.02-0.01}_{+0.02+0.01}$
& $0.01^{-0.00-0.00}_{+0.00+0.00}$
& $A^{D \rho}_2$
& $0.47^{+0.00-0.00}_{-0.00+0.00}$
& $0.89^{+0.00-0.02}_{-0.00+0.02}$
& $0.23^{-0.02-0.01}_{+0.02+0.01}$
\\
$V^{D \omega}$
& $0.85^{-0.02+0.01}_{+0.01-0.01}$
& $1.24^{-0.01-0.00}_{+0.01+0.00}$
& $0.45^{-0.04-0.01}_{+0.04+0.01}$
& $A^{D \omega}_0$
& $0.64^{-0.01+0.01}_{+0.01-0.01}$
& $1.08^{-0.02+0.00}_{+0.01-0.00}$
& $0.50^{-0.04-0.01}_{+0.04+0.01}$
\\
$A^{D \omega}_1$
& $0.58^{-0.01+0.00}_{+0.00-0.00}$
& $0.49^{-0.02-0.01}_{+0.02+0.01}$
& $0.02^{-0.00-0.00}_{+0.01+0.00}$
& $A^{D \omega}_2$
& $0.49^{+0.00-0.00}_{-0.00+0.00}$
& $0.95^{-0.00-0.01}_{-0.00+0.01}$
& $0.28^{-0.02-0.01}_{+0.02+0.01}$
\\
$V^{D K^*}$
& $0.98^{-0.02+0.01}_{+0.02-0.01}$
& $1.10^{-0.02-0.00}_{+0.02+0.00}$
& $0.32^{-0.03-0.01}_{+0.03+0.01}$
& $A^{D K^*}_0$
& $0.78^{-0.01+0.01}_{+0.01-0.01}$
& $1.01^{-0.02-0.00}_{+0.02+0.00}$
& $0.34^{-0.03-0.01}_{+0.03+0.01}$
\\
$A^{D K^*}_1$
& $0.72^{-0.01+0.01}_{+0.01-0.01}$
& $0.45^{-0.02-0.01}_{+0.02+0.01}$
& $0.01^{-0.00-0.00}_{+0.00+0.00}$
& $A^{D K^*}_2$
& $0.60^{+0.00-0.00}_{-0.00+0.00}$
& $0.89^{-0.01-0.01}_{+0.00+0.01}$
& $0.21^{-0.02-0.01}_{+0.02+0.01}$
\\
\hline
$V^{D_s K^*}$
& $0.87^{-0.01+0.01}_{+0.01-0.01}$
& $1.13^{+0.00+0.00}_{-0.01-0.00}$
& $0.69^{-0.04-0.02}_{+0.05+0.03}$
& $A^{D_s K^*}_0$
& $0.61^{-0.00+0.01}_{+0.00-0.01}$
& $0.90^{+0.01+0.01}_{-0.01+0.02}$
& $0.87^{-0.04-0.03}_{+0.05+0.01}$
\\
$A^{D_s K^*}_1$
& $0.56^{-0.00+0.01}_{+0.00-0.01}$
& $0.59^{-0.01-0.01}_{+0.01+0.01}$
& $0.08^{-0.01-0.00}_{+0.01+0.01}$
& $A^{D_s K^*}_2$
& $0.46^{+0.00-0.00}_{-0.00+0.00}$
& $0.90^{+0.01-0.01}_{-0.01+0.01}$
& $0.43^{-0.02-0.02}_{+0.02+0.02}$
\\
$V^{D_s \phi}$
& $0.98^{-0.01+0.00}_{+0.01-0.00}$
& $1.04^{-0.00+0.00}_{+0.00-0.00}$
& $0.54^{-0.03-0.00}_{+0.04+0.01}$
& $A^{D_s \phi}_0$
& $0.72^{-0.01+0.00}_{+0.01-0.00}$
& $0.92^{-0.00+0.00}_{-0.00-0.00}$
& $0.62^{-0.03-0.00}_{+0.04+0.01}$
\\
$A^{D_s \phi}_1$
& $0.69^{-0.00+0.00}_{+0.00-0.00}$
& $0.56^{-0.02-0.00}_{+0.02+0.00}$
& $0.07^{-0.01-0.00}_{+0.01+0.00}$
& $A^{D_s \phi}_2$
& $0.59^{+0.00-0.00}_{-0.00+0.00}$
& $0.90^{-0.00-0.00}_{-0.00+0.00}$
& $0.38^{-0.02-0.00}_{+0.02+0.00}$
\\
\hline
$V^{B \rho}$
& $0.29^{-0.00+0.01}_{-0.00-0.01}$
& $1.77^{-0.03-0.01}_{+0.03+0.01}$
& $1.06^{-0.06-0.03}_{+0.07+0.03}$
& $A^{B \rho}_0$
& $0.32^{+0.00+0.01}_{-0.00-0.01}$
& $1.67^{-0.03-0.01}_{+0.03+0.03}$
& $1.01^{-0.04-0.02}_{+0.05-0.02}$
\\
$A^{B \rho}_1$
& $0.24^{+0.00+0.00}_{-0.00-0.00}$
& $0.86^{-0.03-0.01}_{+0.03+0.01}$
& $0.15^{-0.02-0.01}_{+0.02+0.01}$
& $A^{B \rho}_2$
& $0.22^{+0.00+0.00}_{-0.00-0.00}$
& $1.56^{-0.02-0.02}_{+0.02+0.02}$
& $0.85^{-0.05-0.03}_{+0.05+0.03}$
\\
$V^{B \omega}$
& $0.27^{+0.00+0.00}_{-0.00-0.00}$
& $1.81^{-0.03-0.01}_{+0.03+0.01}$
& $1.18^{-0.07-0.02}_{+0.08+0.02}$
& $A^{B \omega}_0$
& $0.28^{+0.00+0.01}_{-0.00-0.01}$
& $1.62^{+0.05+0.07}_{+0.11+0.09}$
& $1.22^{-0.15-0.11}_{-0.03-0.08}$
\\
$A^{B \omega}_1$
& $0.23^{+0.00+0.00}_{-0.00-0.00}$
& $0.91^{-0.03-0.01}_{+0.03+0.01}$
& $0.18^{-0.02-0.01}_{+0.02+0.01}$
& $A^{B \omega}_2$
& $0.21^{+0.00+0.00}_{-0.00-0.00}$
& $1.62^{-0.03-0.01}_{+0.03+0.01}$
& $0.97^{-0.06-0.03}_{+0.06+0.02}$
\\
$V^{B K^*}$
& $0.36^{-0.00+0.01}_{+0.00-0.01}$
& $1.69^{-0.03-0.01}_{+0.03+0.01}$
& $0.95^{-0.06-0.02}_{+0.06+0.02}$
& $A^{B K^*}_0$
& $0.38^{+0.00+0.01}_{-0.00-0.01}$
& $1.61^{-0.03-0.01}_{+0.03+0.01}$
& $0.89^{-0.04-0.02}_{+0.05+0.02}$
\\
$A^{B K^*}_1$
& $0.31^{+0.00+0.00}_{-0.00-0.00}$
& $0.84^{-0.03-0.01}_{+0.03+0.01}$
& $0.12^{-0.02-0.01}_{+0.02+0.01}$
& $A^{B K^*}_2$
& $0.28^{+0.00+0.00}_{-0.00-0.00}$
& $1.53^{-0.02-0.01}_{+0.02+0.01}$
& $0.79^{-0.04-0.02}_{+0.05+0.02}$
\\
$V^{B D^*}$
& $0.77^{-0.01+0.02}_{+0.00-0.03}$
& $1.25^{-0.02-0.03}_{+0.02+0.02}$
& $0.38^{-0.02-0.03}_{+0.02+0.03}$
& $A^{B D^*}_0$
& $0.68^{-0.00+0.04}_{+0.00-0.04}$
& $1.21^{-0.02-0.03}_{+0.02+0.02}$
& $0.36^{-0.02-0.03}_{+0.02+0.03}$
\\
$A^{B D^*}_1$
& $0.65^{-0.00+0.02}_{+0.00-0.02}$
& $0.60^{-0.01-0.03}_{+0.01+0.02}$
& $0.00^{-0.00-0.01}_{+0.01+0.01}$
& $A^{B D^*}_2$
& $0.61^{-0.00-0.01}_{-0.00-0.00}$
& $1.12^{-0.01-0.05}_{+0.01+0.04}$
& $0.31^{-0.01-0.04}_{+0.01+0.04}$
\\
\hline
$V^{B_s K^*}$
& $0.23^{+0.00+0.01}_{-0.00-0.01}$
& $2.03^{-0.04-0.01}_{+0.04+0.01}$
& $2.27^{-0.20-0.07}_{+0.22+0.08}$
& $A^{B_s K^*}_0$
& $0.25^{+0.00+0.01}_{-0.00-0.01}$
& $1.95^{-0.04-0.01}_{+0.04+0.01}$
& $2.20^{-0.16-0.07}_{+0.18+0.08}$
\\
$A^{B_s K^*}_1$
& $0.19^{+0.00+0.00}_{-0.00-0.01}$
& $1.24^{-0.05-0.02}_{+0.05+0.02}$
& $0.62^{-0.07-0.03}_{+0.09+0.03}$
& $A^{B_s K^*}_2$
& $0.16^{+0.00+0.00}_{-0.00-0.00}$
& $1.83^{-0.04-0.02}_{+0.04+0.02}$
& $1.85^{-0.15-0.07}_{+0.17+0.08}$
\\
$V^{B_s \phi}$
& $0.29^{+0.00+0.00}_{-0.00-0.00}$
& $1.95^{-0.04-0.00}_{+0.04+0.00}$
& $1.98^{-0.17-0.02}_{+0.19+0.02}$
& $A^{B_s \phi}_0$
& $0.31^{+0.00+0.00}_{-0.00-0.00}$
& $1.87^{+0.02-0.00}_{+0.04+0.00}$
& $1.87^{-0.31-0.02}_{+0.16+0.02}$
\\
$A^{B_s \phi}_1$
& $0.25^{+0.00+0.00}_{-0.01-0.00}$
& $1.20^{-0.05-0.01}_{+0.05+0.01}$
& $0.54^{-0.06-0.01}_{+0.07+0.01}$
& $A^{B_s \phi}_2$
& $0.22^{+0.00+0.00}_{-0.01-0.00}$
& $1.79^{-0.04-0.00}_{+0.04+0.00}$
& $1.67^{-0.13-0.02}_{+0.15+0.02}$
\\
$V^{B_s D^*_s}$
& $0.75^{-0.00+0.03}_{+0.00-0.04}$
& $1.37^{-0.03-0.05}_{+0.03+0.04}$
& $0.67^{-0.05-0.08}_{+0.05+0.10}$
& $A^{B_s D^*_s}_0$
& $0.66^{-0.00+0.04}_{-0.00-0.05}$
& $1.33^{-0.03-0.04}_{+0.03+0.04}$
& $0.63^{-0.04-0.08}_{+0.05+0.10}$
\\
$A^{B_s D^*_s}_1$
& $0.62^{+0.00+0.02}_{-0.00-0.03}$
& $0.76^{-0.03-0.05}_{+0.03+0.05}$
& $0.13^{-0.02-0.03}_{+0.02+0.04}$
& $A^{B_s D^*_s}_2$
& $0.57^{+0.00+0.00}_{-0.00-0.01}$
& $1.25^{-0.02-0.07}_{+0.02+0.07}$
& $0.56^{-0.04-0.09}_{+0.04+0.11}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
\subsection{ $P(0^-) \to S (0^+)$ Form Factors}
\begin{itemize}
\item It has been discussed in Sec. III that there are two sets of scalar mesons: the
light scalar nonet formed by $\sigma(600)$,
$\kappa(800)$, $f_0(980)$ and $a_0(980)$; and the heavy scalar nonet
contains $a_0(1450)$, $K_0^*(1430)$, $f_0(1370)$ and
$f_0(1500)/f_0(1710)$. Though their underlying quark structure
is still controversial, the present experimental data seem
to provide a consistent picture that light scalar mesons
below or near 1 GeV can be described by the $qq \overline{q} \overline{q}$ states,
while scalars above 1 GeV form a conventional $q \overline{q}$ with possible
mixing with glueball states. In this work, we have
calculated the form factors involving heavy scalar mesons,
taking $f_0(1710)$ to be primarily a glueball, and $f_0(1500)$ and $f_0(1370)$
to be the SU(3) states as described in Sec. III.
\item From Table~\ref{tab:P2S}, we notice that all the form factors
for charmed mesons are around 0.5-0.6 where as all the
bottom meson form factors lie 0.25 to 0.30, and thus are roughly half of the
charmed meson form factors. Particularly, we note the following patterns:
$F^{B_s K^*_0} = F^{B a_0} = F^{B f_{0q}}$,
$F^{B_s f_{0s}} \approx F^{B D^*_0} = F^{B K^*_0}$, and
$F^{D_s f_{0s}}\approx F^{D a_0} = F^{D f_{0q}}$, where
$f_{0q}= (u\overline{u} + d\overline{d})/\sqrt{2}$, and
$f_{0s}$ is a pure $(s\overline{s})$ state. Here too, the slope parameters
show considerable differences among for the related form factors.
\item Both of the form factors $F^{PS}_{1,0}(0)$ decrease (increase)
with increasing (decreasing) $\beta$ for initial meson,
whereas they increase (decrease) with increasing (decreasing) $\beta$ for final meson.
\item The slope parameters $b$ for both the form factors $F^{PS}_{1,0}$, and
$a$ for $F^{PS}_1$ are found to be positive. For $F^{PS}_0$ form factor, $a$ turns out to be negative
when charmed mesons appear in either initial or final state.
\item For the bottom sector, the slope parameters are larger than that for the charm sector.
The parameter $b$ is generally much small ($< 0.1$) for $F^{PS}_0(0)$, except for
$F^{B_s \to S(0^+)}_0$ and $F^{BD^*}_0$, for which $b$ could be as big as 0.45.
\item For transitions of the charmed mesons, the slope parameters
for the form factor $F^{PS}_0$ are more sensitive to change
in $\beta$ for each of the initial and final mesons. However, these are
less sensitive for $F^{PS}_1$.
\item Slope parameters (except for the case of negative $a$)
show an increase (decrease) with decrease (increase) in $\beta$ for each
of the initial and final mesons.
\item No significant change is found in the form factors, though $a$ and $b$
are slightly lowered that their values obtained in the earlier work \cite{Cheng04}.
Based on the light-cone sum rules, Chernyak \cite{Chernyak} has
estimated the $F^{B a_0(1450)}_{1,0}(0)$ =0.46, while our result is 0.25
and is similar to the $B\to \pi$ form factor at $q^2=0$.
\item On comparison of the $P \to S $ and $P \to P $ form factors,
we notice $F^{D\to S} < F^{D \to P}$ for the same flavor
content of the final state mesons. For the bottom sector,
$F^{B D^*_0} < F^{B D}$ and
$F^{B_s D^*_{s0}} < F^{B_s D_s}$, for heavy-to-heavy transitions, while
$F^{B \to S} \approx F^{B \to P}$ for heavy-to-light transitions.
It has been pointed out before \cite{Cheng04} that the suppression of
the $ B \to D^*_0$ form factor relative to that of $B \to D$ is supported by experiment.
\item An inspection of Table~\ref{tab:P2S}
indicates that similar to the $P \to P $ transitions, the
form factors $F^{PS}_0$ generally show a monopole behavior, and $F^{PS}_1$
have a dipole behavior particularly for charmed meson
transitions. In general, form factors for $P \to S $
transitions increase slowly with $q^2$ compared to that for $P \to P $ ones.
\end {itemize}
{\squeezetable
\begin{table}[b]
\caption{Form factors of
$P(0^-)\rightarrow S(0^+)$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless. }
\label{tab:P2S}
\begin{ruledtabular}
\begin{tabular}{|cccc|cccc|}
$F$
& $F(0)$
& $a$
& $b$
& $F$
& $F(0)$
& $a$
& $b$
\\
\hline
$F^{D a_0}_1$
& $0.51^{-0.01+0.01}_{+0.01-0.02}$
& $1.06^{-0.02-0.02}_{+0.01+0.01}$
& $0.24^{-0.02-0.02}_{+0.02+0.02}$
& $F^{D a_0}_0$
& $0.51^{-0.01+0.02}_{+0.01-0.02}$
& $-0.04^{+0.07-0.06}_{-0.06+0.07}$
& $0.02^{-0.02+0.01}_{+0.02-0.02}$
\\
$F^{D f_{0q}}_1$
& $0.51^{-0.01+0.03}_{+0.01-0.04}$
& $1.06^{-0.02-0.04}_{+0.01+0.03}$
& $0.24^{-0.02-0.04}_{+0.02+0.04}$
& $F^{D f_{0q}}_0$
& $0.51^{-0.01+0.04}_{+0.01-0.05}$
& $-0.04^{+0.07-0.13}_{-0.06+0.19}$
& $0.02^{-0.02+0.03}_{+0.02-0.04}$
\\
$F^{DK^*_0}_1$
& $0.47^{-0.01+0.02}_{+0.01-0.02}$
& $0.94^{-0.02-0.01}_{+0.02+0.01}$
& $0.19^{-0.02-0.01}_{+0.02+0.02}$
& $F^{DK^*_0}_0$
& $0.47^{-0.01+0.02}_{+0.01-0.03}$
& $-0.31^{+0.04-0.04}_{-0.04+0.05}$
& $0.08^{-0.01+0.01}_{+0.01-0.01}$
\\
\hline
$F^{D_s K^*_0}_1$
& $0.55^{-0.01+0.02}_{+0.01-0.03}$
& $1.02^{-0.01-0.01}_{+0.00+0.01}$
& $0.38^{-0.02-0.04}_{+0.02+0.06}$
& $F^{D_s K^*_0}_0$
& $0.55^{-0.01+0.02}_{+0.01-0.03}$
& $-0.04^{+0.02-0.05}_{-0.02+0.07}$
& $0.05^{-0.01+0.01}_{+0.01-0.01}$
\\
$F^{D_s f_{0s}}_1$
& $0.52^{-0.01+0.01}_{+0.01-0.01}$
& $0.91^{-0.01-0.00}_{+0.01+0.00}$
& $0.29^{-0.02-0.01}_{+0.02+0.01}$
& $F^{D_s f_{0s}}_0$
& $0.52^{-0.01+0.01}_{+0.01-0.01}$
& $-0.34^{+0.01-0.01}_{-0.01+0.01}$
& $0.10^{-0.01+0.00}_{+0.01-0.00}$
\\
\hline
$F^{B a_0}_1$
& $0.25^{-0.00+0.01}_{+0.00-0.01}$
& $1.53^{-0.03-0.01}_{+0.03+0.01}$
& $0.64^{-0.04-0.04}_{+0.05+0.04}$
& $F^{B a_0}_0$
& $0.25^{-0.00+0.01}_{+0.00-0.01}$
& $0.54^{-0.01-0.03}_{+0.01+0.03}$
& $0.01^{-0.01-0.00}_{+0.02+0.00}$
\\
$F^{B f_{0q}}_1$
& $0.25^{-0.00+0.03}_{+0.00-0.03}$
& $1.53^{-0.03-0.03}_{+0.03+0.03}$
& $0.64^{-0.04-0.08}_{+0.05+0.11}$
& $F^{B f_{0q}}_0$
& $0.25^{-0.00+0.03}_{+0.00-0.03}$
& $0.54^{-0.01-0.06}_{+0.01+0.07}$
& $0.01^{-0.01-0.00}_{+0.02+0.01}$
\\
$F^{BK^*_0}_1$
& $0.27^{-0.01+0.01}_{+0.01-0.02}$
& $1.43^{-0.03-0.01}_{+0.03+0.01}$
& $0.52^{-0.04-0.03}_{+0.04+0.04}$
& $F^{BK^*_0}_0$
& $0.27^{-0.01+0.01}_{+0.01-0.02}$
& $0.32^{-0.01-0.02}_{+0.01+0.03}$
& $0.05^{-0.01+0.00}_{+0.01+0.00}$
\\
$F^{BD^*_0}_1$
& $0.27^{-0.01+0.03}_{+0.01-0.03}$
& $1.08^{-0.04+0.03}_{+0.04-0.07}$
& $0.23^{-0.02-0.00}_{+0.02+0.00}$
& $F^{B D^*_0}_0$
& $0.27^{-0.01+0.03}_{+0.01-0.03}$
& $-0.48^{+0.02-0.03}_{-0.02+0.01}$
& $0.36^{-0.03+0.03}_{+0.03-0.03}$
\\
\hline
$F^{B_s K^*_0}_1$
& $0.25^{-0.00+0.02}_{+0.00-0.02}$
& $1.75^{-0.04-0.04}_{+0.04+0.05}$
& $1.33^{-0.12-0.12}_{+0.14+0.18}$
& $F^{B_s K^*_0}_0$
& $0.25^{-0.00+0.02}_{+0.00-0.02}$
& $0.74^{-0.03-0.06}_{+0.03+0.07}$
& $0.24^{-0.04-0.03}_{+0.04+0.05}$
\\
$F^{B_s f_{0s}}_1$
& $0.28^{-0.00+0.01}_{+0.00-0.01}$
& $1.64^{-0.04-0.01}_{+0.04+0.01}$
& $1.07^{-0.10-0.04}_{+0.11+0.04}$
& $F^{B_s f_{0s}}_0$
& $0.28^{-0.00+0.01}_{+0.00-0.01}$
& $0.52^{-0.03-0.02}_{+0.03+0.02}$
& $0.20^{-0.03-0.01}_{+0.03+0.01}$
\\
$F^{B_s D^*_{0s}}_1$
& $0.30^{-0.02+0.03}_{+0.02-0.03}$
& $1.18^{-0.06+0.01}_{+0.06-0.04}$
& $0.51^{-0.05-0.05}_{+0.06+0.05}$
& $F^{B_s D^*_{0s}}_0$
& $0.30^{-0.02+0.03}_{+0.02-0.04}$
& $-0.47^{-0.01-0.02}_{+0.01+0.01}$
& $0.45^{-0.02+0.01}_{+0.02+0.00}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
}
\subsection{$P(0^-) \to A (1^+\,:~ ^3P_1)$ Form Factors}
\begin{itemize}
\item In Tables~\ref{tab:P2Ap} and \ref{tab:P2An}, we have given heavy-to-light
form factors involving axial vector nonet mesons, whereas
heavy to heavy form factors are separately presented in
Table ~\ref{tab:B2D} with the final state charmed mesons
being taken as the heavy quark spin basis.
\item From Table~\ref{tab:P2Ap}, all the form factors are found to be positive
for the bottom as well as charm sectors. We also notice the following pattern:
$V^{PA}_1 > V^{PA}_0 > A^{PA} > V^{PA}_2$ for the charmed mesons
and $V^{PA}_1 > A^{PA} > V^{PA}_2 > V^{PA}_0$ for the bottom mesons.
\item Numerically speaking, the form factor $A^{PA}(0)$ for the bottom transitions
is generally around 0.25, and it is larger
than that for charmed meson transition for which it lies close 0.16. Similar
behavior is observed for $V^{PA}_2(0)$, which is $<0.1$ for the charm sector, where as
it lies around 0.2 for the bottom transitions. In contrast, the form factor
$V^{PA}_1(0)$, lying around 0.4 for the bottom sector, is significantly smaller than that
for the charm sector, where its value lies between 1.4 to 1.8. Also $V^{PA}_0(0)$ for the
bottom transitions is roughly half of its value for charmed meson
transitions.
\item The form factors are not very sensitive to the variation chosen for
the $\beta$ parameters. However, they generally
tend to decrease (increase) with increasing (decreasing) $\beta$ for initial meson, whereas
they increase (decrease) with increasing (decreasing) $\beta$ for final meson.
\item All the slope parameters, except for $V^{PA}_1$ and $V^{PA}_2$ for the charmed meson
transitions, are found to be positive as per the definition given in
Eq.~(\ref{eq:FFpara}). For the bottom sector, the slope parameters
are significantly larger than that for the charm sector.
\item Slope parameters $a$ and $b$ for $A^{PA}$, and $V^{PA}_1$ are
generally less sensitive to $\beta$ variation, but for $V^{PA}_0$ and $V^{PA}_2$ they
could show more sensitivity (even up to $20\%$) to change in $\beta$ values.
\item Generally no large change occurs in the form factors obtained in
the earlier work \cite{Cheng04}, though slope
parameters often show some changes.\footnote{Form factor $A^{D K_1(^3P_1)}(0) = 0.98$ given
in the earlier \cite{Cheng04} is erroneous, and should be replaced
with 0.15.} This could happen since we now perform 5-point fit for $q^2$ values.
We find that the form factors $V^{PA}_0$
for $B \to a_1$ transition has marginally increased to 0.14.
\item There are several existing model calculations for
$B\rightarrow A$ form factors: the ISGW2 model \cite{ISGW2},
the constituent quark-meson model (CQM) \cite{Deandrea},
the QCD sum rules (QSR) \cite{Aliev}, light cone sum rules
(LCSR) \cite{KCYang08}, and more recently the
perturbative QCD (pQCD) approach \cite{RHLi09}. For the sake of comparison, results
for $B \to a_1$ transition form factors are given in Table \ref{tab:B2a1} for these
approaches, which show quite significant differences since these approaches differ in
their treatment of dynamics of form factors. For
example, $V^{Ba_1}_0 = 1.20$, obtained in the quark-meson model and 1.01 in the ISGW2 model,
are larger than the values obtained in other approaches. If $a_1 (1260)$
behaves as the scalar partner of the $\rho$ meson, $V^{Ba_1}_0$ is expected to
be similar to $A^{B\rho}_0$, which is of order 0.3 at $q^2 = 0$. Therefore, it
appears to us that a magnitude of order unity for $V^{Ba_1}_0 $ as predicted
by the ISGW2 model and CQM is very unlikely. In principle, the experimental measurements
of $\overline{B}^0\rightarrow a^{\pm}_1 \pi^{\mp}$ will enable us to test the
form factors $V^{Ba_1}_0$. The BaBar and Belle measurements \cite{BABAR, Abe07} of
$\overline{B}^0\rightarrow a^{\pm}_1 \pi^{\mp}$ favors a value of
$V_0^{Ba_1}(0)\approx 0.30$ \cite{ChengAP}, which is very close the LCSR result
shown in Table \ref{tab:B2a1}.
\end {itemize}
{\squeezetable
\begin{table}[b]
\caption{Form factors of
$P(0^-)\rightarrow A(1^{++})$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless. }
\label{tab:P2Ap}
\begin{ruledtabular}
\begin{tabular}{| c c c c | c c c c |}
$F$
& $F(0)$
& $a$
& $b$
& $F$
& $F(0)$
& $a$
& $b$
\\
\hline
$A^{D a_1}$
& $0.19^{-0.01+0.00}_{+0.01-0.01}$
& $1.03^{-0.03-0.01}_{+0.02+0.00}$
& $0.16^{-0.02-0.01}_{+0.02+0.01}$
& $V^{D a_1}_0$
& $0.32^{-0.00-0.00}_{+0.00-0.00}$
& $0.96^{-0.01-0.01}_{+0.00+0.00}$
& $0.43^{-0.06+0.01}_{+0.07-0.01}$
\\
$V^{D a_1}_1$
& $1.51^{-0.04+0.00}_{+0.04-0.01}$
& $-0.06^{-0.01-0.02}_{+0.01+0.02}$
& $0.04^{-0.00+0.00}_{+0.00-0.00}$
& $V^{D a_1}_2$
& $0.05^{-0.01+0.00}_{+0.01-0.00}$
& $-0.02^{-0.08-0.00}_{+0.07+0.00}$
& $0.12^{+0.00-0.01}_{-0.00+0.01}$
\\
$A^{D f_{1q}}$
& $0.18^{-0.01+0.01}_{+0.01-0.01}$
& $1.03^{-0.03-0.02}_{+0.02+0.00}$
& $0.16^{-0.02-0.02}_{+0.02+0.03}$
& $V^{D f_{1q}}_0$
& $0.34^{-0.00-0.00}_{+0.00-0.00}$
& $0.97^{-0.02-0.01}_{+0.01-0.00}$
& $0.39^{-0.05+0.03}_{+0.06+0.00}$
\\
$V^{D f_{1q}}_1$
& $1.75^{-0.05-0.00}_{+0.04-0.02}$
& $-0.02^{-0.01-0.05}_{+0.01+0.06}$
& $0.04^{-0.00+0.01}_{+0.00-0.00}$
& $V^{D f_{1q}}_2$
& $0.05^{-0.01+0.00}_{+0.01-0.00}$
& $-0.02^{-0.08-0.01}_{+0.07+0.01}$
& $0.12^{+0.00-0.01}_{-0.00+0.02}$
\\
$A^{D K_{1A}}$
& $0.15^{-0.01+0.01}_{+0.01-0.01}$
& $0.89^{-0.03+0.00}_{+0.03-0.01}$
& $0.12^{-0.02-0.01}_{+0.02+0.01}$
& $V^{D K_{1A}}_0$
& $0.28^{-0.00-0.00}_{+0.00+0.00}$
& $0.84^{-0.02-0.01}_{+0.01-0.01}$
& $0.39^{-0.05+0.04}_{+0.06-0.03}$
\\
$V^{D K_{1A}}_1$
& $1.60^{-0.05-0.02}_{+0.05+0.01}$
& $-0.22^{-0.00-0.03}_{+0.00+0.03}$
& $0.07^{-0.00+0.00}_{+0.00-0.00}$
& $V^{D K_{1A}}_2$
& $0.01^{-0.00+0.00}_{+0.00-0.00}$
& $-0.83^{-0.17+0.02}_{+0.15-0.03}$
& $0.24^{+0.04-0.01}_{-0.03+0.01}$
\\
\hline
$A^{D_s K_{1A}}$
& $0.19^{-0.01+0.01}_{+0.01-0.01}$
& $0.99^{-0.01-0.01}_{+0.01+0.01}$
& $0.28^{-0.02-0.03}_{+0.02+0.04}$
& $V^{D_s K_{1A}}_0$
& $0.29^{-0.00-0.00}_{+0.00+0.00}$
& $0.72^{+0.05-0.09}_{-0.06+0.07}$
& $0.87^{-0.09+0.08}_{+0.10-0.05}$
\\
$V^{D_s K_{1A}}_1$
& $1.68^{-0.03-0.01}_{+0.03+0.00}$
& $-0.04^{-0.01-0.04}_{+0.01+0.05}$
& $0.06^{-0.00+0.00}_{+0.00-0.00}$
& $V^{D_s K_{1A}}_2$
& $0.07^{-0.00-0.00}_{+0.00-0.00}$
& $0.22^{-0.03-0.01}_{+0.03+0.02}$
& $0.15^{-0.01-0.01}_{+0.01+0.02}$
\\
$A^{D_s f_{1s}}$
& $0.17^{-0.01+0.00}_{+0.01-0.00}$
& $0.86^{-0.02-0.00}_{+0.02-0.00}$
& $0.20^{-0.02-0.01}_{+0.02+0.01}$
& $V^{D_s f_{1s}}_0$
& $0.22^{+0.00-0.00}_{-0.00+0.00}$
& $0.19^{+0.13-0.13}_{-0.17+0.10}$
& $1.20^{-0.15+0.11}_{+0.18-0.09}$
\\
$V^{D_s f_{1s}}_1$
& $1.47^{-0.03-0.01}_{+0.03+0.01}$
& $-0.29^{-0.01-0.01}_{+0.01+0.01}$
& $0.09^{-0.00+0.00}_{+0.00-0.00}$
& $V^{D_s f_{1s}}_2$
& $0.03^{-0.00+0.00}_{+0.00-0.00}$
& $-0.34^{-0.05+0.00}_{+0.05-0.00}$
& $0.17^{+0.00-0.00}_{-0.00+0.00}$
\\
\hline
$A^{B a_1}$
& $0.24^{-0.01+0.01}_{+0.01-0.01}$
& $1.48^{-0.03-0.01}_{+0.03+0.01}$
& $0.57^{-0.04-0.03}_{+0.05+0.04}$
& $V^{B a_1}_0$
& $0.14^{+0.01+0.01}_{-0.01-0.01}$
& $1.66^{-0.04-0.01}_{+0.04+0.01}$
& $1.11^{-0.08-0.03}_{+0.09+0.02}$
\\
$V^{B a_1}_1$
& $0.36^{-0.01+0.01}_{+0.01-0.01}$
& $0.26^{-0.02-0.02}_{+0.02+0.03}$
& $0.14^{-0.01-0.00}_{+0.01+0.01}$
& $V^{B a_1}_2$
& $0.17^{-0.01+0.01}_{+0.01-0.01}$
& $1.08^{-0.05-0.02}_{+0.05+0.02}$
& $0.44^{-0.03-0.03}_{+0.09+0.04}$
\\
$A^{B f_{1q}}$
& $0.24^{-0.01+0.02}_{+0.01-0.03}$
& $1.48^{-0.03-0.03}_{+0.03+0.03}$
& $0.57^{-0.04-0.08}_{+0.05+0.10}$
& $V^{B f_{1q}}_0$
& $0.14^{+0.01+0.01}_{-0.01-0.02}$
& $1.65^{-0.04-0.02}_{+0.04+0.02}$
& $1.07^{-0.08-0.06}_{+0.09+0.09}$
\\
$V^{B f_{1q}}_1$
& $0.37^{-0.01+0.03}_{+0.01-0.03}$
& $0.27^{-0.02-0.05}_{+0.02+0.06}$
& $0.13^{-0.01-0.01}_{+0.01+0.02}$
& $V^{B f_{1q}}_2$
& $0.17^{-0.01+0.01}_{+0.01-0.02}$
& $1.08^{-0.05-0.04}_{+0.05+0.05}$
& $0.44^{-0.03-0.07}_{+0.03+0.09}$
\\
$A^{B K_{1A}}$
& $0.27^{-0.01+0.01}_{+0.01-0.02}$
& $1.39^{-0.04-0.01}_{+0.04+0.00}$
& $0.47^{-0.04-0.03}_{+0.04+0.04}$
& $V^{B K_{1A}}_0$
& $0.16^{+0.01+0.01}_{-0.01-0.01}$
& $1.55^{-0.03+0.01}_{+0.04-0.01}$
& $1.00^{-0.10-0.02}_{+0.09+0.03}$
\\
$V^{B K_{1A}}_1$
& $0.39^{-0.01+0.01}_{+0.01-0.02}$
& $0.07^{-0.02-0.02}_{+0.02+0.02}$
& $0.19^{-0.00-0.00}_{+0.00+0.01}$
& $V^{B K_{1A}}_2$
& $0.17^{-0.01+0.01}_{+0.01-0.01}$
& $0.84^{-0.06-0.01}_{+0.06+0.01}$
& $0.36^{-0.02-0.03}_{+0.02+0.04}$
\\
\hline
$A^{B_s K_{1A}}$
& $0.24^{-0.00+0.02}_{+0.00-0.02}$
& $1.70^{-0.05-0.04}_{+0.05+0.05}$
& $1.22^{-0.12-0.12}_{+0.13+0.17}$
& $V^{B_s K_{1A}}_0$
& $0.12^{+0.01+0.01}_{-0.01-0.01}$
& $1.88^{-0.05-0.02}_{+0.06+0.03}$
& $2.06^{-0.25-0.08}_{+0.30+0.13}$
\\
$V^{B_s K_{1A}}_1$
& $0.37^{-0.00+0.02}_{+0.00-0.03}$
& $0.53^{-0.04-0.06}_{+0.04+0.07}$
& $0.29^{-0.03-0.03}_{+0.03+0.04}$
& $V^{B_s K_{1A}}_2$
& $0.17^{-0.00+0.01}_{+0.00-0.01}$
& $1.40^{-0.07-0.04}_{+0.07+0.05}$
& $0.97^{-0.10-0.10}_{+0.12+0.15}$
\\
$A^{B_s f_{1s}}$
& $0.28^{-0.01+0.01}_{+0.01-0.01}$
& $1.59^{-0.05-0.01}_{+0.05+0.01}$
& $0.99^{-0.10-0.03}_{+0.11+0.04}$
& $V^{B_s f_{1s}}_0$
& $0.13^{+0.01+0.00}_{-0.01-0.00}$
& $1.79^{-0.06-0.00}_{+0.05+0.00}$
& $2.00^{-0.25-0.01}_{+0.33+0.02}$
\\
$V^{B_s f_{1s}}_1$
& $0.41^{-0.01+0.01}_{+0.01-0.01}$
& $0.29^{-0.04-0.02}_{+0.04+0.02}$
& $0.29^{-0.02-0.01}_{+0.02+0.01}$
& $V^{B_s f_{1s}}_2$
& $0.18^{-0.01+0.00}_{+0.01-0.00}$
& $1.18^{-0.07-0.01}_{+0.07+0.01}$
& $0.74^{-0.07-0.03}_{+0.08+0.03}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
}
{\squeezetable
\begin{table}[b1]
\caption{Form factors of
$B \rightarrow a_1$ transitions at maximum recoil ($q^2=0$). The results of CQM and
QSR have been rescaled according to the form factor definition in Eq. (\ref{eq:ffpdimless})}
\label{tab:B2a1}
\begin{ruledtabular}
\begin{tabular}{| c c c c c c c |}
$B \rightarrow a_1$
& $\textrm{This work}$
& $\textrm{ISGW2}$ \cite{ISGW2}
& $\textrm{CQM}$ \cite{Deandrea}
& $\textrm{QSR}$ \cite{Aliev}
& $\textrm{LCSR}$ \cite{KCYang08}
& $\textrm{pQCD}$ \cite{RHLi09}
\\
\hline
$A$
& $0.24^{-0.01+0.01}_{+0.01-0.01}$
& $0.21$
& $0.09$
& $0.41\pm 0.06$
& $0.48\pm 0.09$
& $0.26^{+0.06+0.00+0.03}_{-0.05-0.01-0.03}$
\\
$V_0$
& $0.14^{+0.01+0.01}_{-0.01-0.01}$
& $1.01$
& $1.20$
& $0.23\pm 0.05$
& $0.30\pm 0.05$
& $0.34^{+0.07+0.01+0.08}_{-0.07-0.02-0.08}$
\\
$V_1$
& $0.36^{-0.01+0.01}_{+0.01-0.01}$
& $0.54$
& $1.32$
& $0.68\pm 0.08$
& $0.37\pm 0.07$
& $0.43^{+0.10+0.01+0.05}_{-0.09-0.01-0.05}$
\\
$V_2$
& $0.17^{-0.01+0.01}_{+0.01-0.01}$
& $-0.05$
& $0.34$
& $0.33\pm 0.03$
& $0.42\pm 0.08$
& $0.13^{+0.03+0.00+0.00}_{-0.03-0.01-0.00}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
}
\subsection{ $ P(0^-)\to A (1^+\,:~ ^1P_1 )$ Form Factors}
\begin{itemize}
\item From Table~\ref{tab:P2An}, we find that the form factors,
$A^{PA}(0), V^{PA}_0(0)$ and $V^{PA}_1(0)$, are positive,
where as $V^{PA}_2(0)$ is negative and small (around -0.1) for the
bottom as well as charm sector, and follow the pattern:
$V^{PA}_1 > V^{PA}_0 > A^{PA} > |V^{PA}_2| $ for the charmed mesons
and $V^{PA}_0 > V^{PA}_1 > A^{PA} \geq |V^{PA}_2| $ for the bottom mesons.
Numerically, form factors $A^{PA}(0)$ is generally
around 0.1, where as $V^{PA}_0(0)$ lies close to 0.5 for all the cases.
The form factor $V^{PA}_1(0)$, lying between 0.15 to 0.20, for the
bottom sector is significantly smaller than that
for the charm sector, where its value lies between 1.3 to 1.6.
\item The form factors
$A^{PA}_0(0)$ and $V^{PA}_1(0)$ usually increase (decrease) with
increasing (decreasing) beta for initial meson as well as for final meson.
\item The form factors $V^{PA}_0(0)$ and $V^{PA}_3(0)$ decrease (increase) in magnitude
with increasing (decreasing) $\beta$ for initial meson, and show the opposite
trend for final mesons, i.e., these increase (decrease) with
increasing (decreasing) $\beta$ for the final state.
\item All the slope parameters are found to be positive. For the bottom sector,
the slope parameters are larger than that for the charm sector.
\item Slope parameters $a$ and $b$ for $A^{PA}, V^{PA}_0$, and $V^{PA}_1$
are less sensitive (a few $\%$) to the variation in the $\beta$ values. For
$V^{PA}_2$ form factor, the slope parameters show huge sensitivity to the change in beta values,
even with assuming $q^2$ behavior given by Eq. (\ref{eq:FFpara}).
However, these are less sensitive for $B \to b_1/h_{1q}$ cases.
\item No significant change is found in the form factors obtained in
the earlier work \cite{Cheng04}, however, the slope parameters show
difference.
\item While comparing the form factors of heavy-to-light spin $1$ meson transitions,
we notice the following relations for the same flavor content of the mesons:
$A^{P A(1^{++})}(0) > A^{P A(1^{+-})}(0)~$,
$V^{P A(1^{++})}_1(0) > V^{P A(1^{+-})}_1(0)$. But for $V^{PA}_0(0)$ form factors, we find
$V^{P A(1^{++})}_0(0) < V^{P A(1^{+-})}_0(0)$. For $V^{PA}_2(0)$ form factors, we observe
opposite behavior for the charmed and bottom mesons, i.e.,
$|V^{D, D_s\to A(1^{++})}_2(0)| < |V^{D, Ds \to A(1^{+-})}_2(0)|$,
whereas $|V^{B, Bs \to A(1^{++})}_2(0)| > |V^{B, Bs \to A(1^{+-})}_2(0)|$.
\item While comparing the form factors of heavy-to-light vector and axial-vector mesons,
we notice the following patterns:
$V^{P V}(0) > A^{P A(1^{++})}(0)$ for the same flavor content of the mesons;
$ V^{P A(1^{+-})}_1(0)> A^{P V}_1(0)$ for charmed mesons
and $A^{P V}_1(0) > V^{P A(1^{+-})}_1(0)$ for the bottom mesons.
But for $V^{PA}_0(0)$ form factors, we find
$V^{P A(1^{+-})}_0(0)< A^{P V}_0(0)$ for charmed mesons
and $V^{P A(1^{++})}_0(0)< A^{P V}_0(0) < V^{P A(1^{+-})}_0(0)$ for the bottom mesons.
We also observe that $A^{P V}_2(0)$ is higher than both $V^{D, Ds \to A(1^{++})}_2(0)$
as well as $|V^{D, Ds \to A(1^{+-})}_2(0)| $.
\end {itemize}
{\squeezetable
\begin{table}[b]
\caption{Form factors of
$P(0^-)\rightarrow A(1^{+-})$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless.}
\label{tab:P2An}
\begin{ruledtabular}
\begin{tabular}{| c c c c | c c c c |}
$F$
& $F(0)$
& $a$
& $b$
& $F$
& $F(0)$
& $a$
& $b$
\\
\hline
$A^{D b_1}$
& $0.12^{+0.00+0.00}_{-0.00-0.02}$
& $1.09^{-0.01-0.01}_{+0.01+0.01}$
& $0.50^{-0.04-0.04}_{+0.04+0.04}$
& $V^{D b_1}_0$
& $0.50^{-0.01+0.02}_{+0.01-0.02}$
& $0.98^{-0.02-0.01}_{+0.02-0.00}$
& $0.26^{-0.00-0.02}_{+0.00+0.03}$
\\
$V^{D b_1}_1$
& $1.39^{+0.02+0.03}_{-0.02-0.04}$
& $0.44^{-0.03-0.02}_{+0.03+0.02}$
& $0.05^{-0.01-0.00}_{+0.01+0.01}$
& $V^{D b_1}_2$
& $-0.10^{+0.02-0.01}_{-0.02+0.01}$
& $0.26^{-0.64+0.13}_{+0.28-0.23}$
& $0.90^{+0.46-0.16}_{-0.23+0.23}$
\\
$A^{D h_{1q}}$
& $0.11^{+0.00+0.00}_{-0.00-0.01}$
& $1.09^{-0.01-0.02}_{+0.01+0.01}$
& $0.50^{-0.04-0.08}_{+0.04+0.10}$
& $V^{D h_{1q}}_0$
& $0.49^{-0.01+0.04}_{+0.01-0.05}$
& $0.98^{-0.02-0.03}_{+0.02-0.02}$
& $0.26^{-0.00-0.06}_{+0.00+0.06}$
\\
$V^{D h_{1q}}_1$
& $1.42^{+0.02+0.06}_{-0.02-0.09}$
& $0.44^{-0.03-0.04}_{+0.03+0.04}$
& $0.05^{-0.01-0.01}_{+0.01+0.02}$
& $V^{D h_{1q}}_2$
& $-0.10^{+0.02-0.02}_{-0.01+0.03}$
& $0.26^{-0.64+0.24}_{+0.28-0.79}$
& $0.90^{+0.46-0.33}_{-0.23+0.69}$
\\
$A^{D K_{1B}}$
& $0.10^{+0.00+0.00}_{-0.00-0.00}$
& $0.98^{-0.01-0.01}_{+0.01+0.01}$
& $0.37^{-0.03-0.03}_{+0.03+0.04}$
& $V^{D K_{1B}}_0$
& $0.48^{-0.01+0.02}_{+0.01-0.03}$
& $0.94^{-0.02-0.02}_{+0.01+0.01}$
& $0.22^{+0.00-0.03}_{-0.00+0.03}$
\\
$V^{D K_{1B}}_1$
& $1.58^{+0.02+0.03}_{-0.03-0.05}$
& $0.31^{-0.02-0.01}_{+0.02+0.02}$
& $0.04^{-0.00-0.00}_{+0.00+0.01}$
& $V^{D K_{1B}}_2$
& $-0.13^{+0.01-0.01}_{-0.01+0.01}$
& $0.57^{-0.06-0.01}_{+0.04-0.01}$
& $0.32^{+0.05-0.04}_{-0.03+0.06}$
\\
\hline
$A^{D_s K_{1B}}$
& $0.10^{+0.00+0.00}_{-0.00-0.00}$
& $0.97^{+0.01+0.02}_{+0.01-0.16}$
& $0.71^{-0.04-0.08}_{+0.03+0.19}$
& $V^{D_s K_{1B}}_0$
& $0.51^{-0.01+0.03}_{+0.01-0.04}$
& $0.91^{-0.01+0.00}_{+0.01-0.03}$
& $0.45^{-0.00-0.07}_{+0.00+0.10}$
\\
$V^{D_s K_{1B}}_1$
& $1.50^{+0.01+0.05}_{-0.01-0.07}$
& $0.59^{-0.02-0.02}_{+0.02+0.03}$
& $0.10^{-0.01-0.01}_{+0.01+0.02}$
& $V^{D_s K_{1B}}_2$
& $-0.12^{+0.01-0.01}_{-0.01+0.01}$
& $0.68^{-0.02-0.02}_{+0.01+0.01}$
& $0.36^{+0.01-0.05}_{-0.01+0.08}$
\\
$A^{D_s h_{1s}}$
& $0.10^{+0.00+0.00}_{-0.00-0.00}$
& $0.93^{-0.00-0.00}_{-0.00-0.00}$
& $0.51^{-0.03-0.02}_{+0.03+0.02}$
& $V^{D_s h_{1s}}_0$
& $0.57^{-0.01+0.01}_{+0.01-0.01}$
& $0.89^{-0.01+0.00}_{+0.01-0.00}$
& $0.37^{+0.00-0.02}_{-0.00+0.03}$
\\
$V^{D_s h_{1s}}_1$
& $1.43^{+0.01+0.02}_{-0.01-0.02}$
& $0.46^{-0.02-0.01}_{+0.02+0.01}$
& $0.07^{-0.01-0.00}_{+0.01+0.00}$
& $V^{D_s h_{1s}}_2$
& $-0.17^{+0.01-0.00}_{-0.01+0.00}$
& $0.55^{-0.02-0.01}_{+0.01+0.01}$
& $0.20^{-0.01-0.01}_{+0.00+0.01}$
\\
\hline
$A^{B b_1}$
& $0.11^{+0.00+0.00}_{-0.00-0.01}$
& $1.89^{-0.03-0.03}_{+0.03+0.03}$
& $1.51^{-0.08-0.09}_{+0.09+0.10}$
& $V^{B b_1}_0$
& $0.38^{-0.01+0.03}_{+0.01-0.03}$
& $1.38^{-0.03+0.00}_{+0.03-0.01}$
& $0.63^{-0.02-0.04}_{+0.03+0.04}$
\\
$V^{B b_1}_1$
& $0.19^{+0.01+0.01}_{-0.01-0.01}$
& $0.99^{-0.03-0.03}_{+0.03+0.04}$
& $0.29^{-0.03-0.03}_{+0.03+0.03}$
& $V^{B b_1}_2$
& $-0.02^{+0.01-0.01}_{-0.01+0.00}$
& $1.11^{-20.4+0.33}_{+0.75-0.75}$
& $7.76^{+59.7-1.82}_{-4.34+3.28}$
\\
$A^{B h_{1q}}$
& $0.10^{+0.00+0.01}_{-0.00-0.01}$
& $1.89^{-0.03-0.07}_{+0.03+0.07}$
& $1.51^{-0.08-0.20}_{+0.09+0.25}$
& $V^{B h_{1q}}_0$
& $0.37^{-0.01+0.07}_{+0.01-0.06}$
& $1.37^{-0.03-0.00}_{+0.03-0.03}$
& $0.62^{-0.02-0.08}_{+0.03+0.11}$
\\
$V^{B h_{1q}}_1$
& $0.19^{+0.01+0.02}_{-0.01-0.02}$
& $0.99^{-0.03-0.08}_{+0.03+0.08}$
& $0.29^{-0.03-0.06}_{+0.03+0.09}$
& $V^{B h_{1q}}_2$
& $-0.02^{+0.01-0.01}_{-0.01+0.01}$
& $1.11^{-20.4+0.53}_{+0.75-3.74}$
& $7.76^{+59.7-3.38}_{-4.35+13.7}$
\\
$A^{B K_{1B}}$
& $0.12^{+0.00+0.01}_{-0.00-0.01}$
& $1.78^{-0.03-0.03}_{+0.03+0.04}$
& $1.27^{-0.07-0.08}_{+0.07+0.11}$
& $V^{B K_{1B}}_0$
& $0.45^{-0.01+0.04}_{+0.01-0.04}$
& $1.37^{-0.03-0.00}_{+0.03-0.01}$
& $0.54^{-0.02-0.04}_{+0.02+0.05}$
\\
$V^{B K_{1B}}_1$
& $0.21^{+0.01+0.01}_{-0.01-0.01}$
& $0.83^{-0.03-0.03}_{+0.03+0.04}$
& $0.22^{-0.02-0.02}_{+0.02+0.03}$
& $V^{B K_{1B}}_2$
& $-0.05^{+0.01-0.01}_{-0.01+0.01}$
& $1.74^{-0.08-0.03}_{-0.00+0.02}$
& $2.17^{-1.09-0.23}_{-0.53+0.31}$
\\
\hline
$A^{B_s K_{1B}}$
& $0.08^{+0.00+0.01}_{-0.00-0.01}$
& $2.06^{-0.04-0.04}_{+0.03+0.05}$
& $2.57^{-0.20-0.23}_{+0.23+0.33}$
& $V^{B_s K_{1B}}_0$
& $0.38^{-0.01+0.04}_{+0.01-0.05}$
& $1.64^{-0.04-0.03}_{+0.04+0.04}$
& $1.25^{-0.07-0.13}_{+0.09+0.19}$
\\
$V^{B_s K_{1B}}_1$
& $0.15^{+0.01+0.01}_{-0.01-0.02}$
& $1.34^{-0.05-0.06}_{+0.05+0.07}$
& $0.76^{-0.08-0.10}_{+0.10+0.14}$
& $V^{B_s K_{1B}}_2$
& $-0.06^{+0.01-0.01}_{-0.01+0.01}$
& $1.65^{-0.01-0.03}_{+0.02+0.04}$
& $1.16^{-0.02-0.09}_{+0.06+0.13}$
\\
$A^{B_s h_{1s}}$
& $0.09^{+0.00+0.00}_{-0.00-0.00}$
& $1.95^{-0.04-0.01}_{+0.04+0.01}$
& $2.11^{-0.16-0.07}_{+0.18+0.07}$
& $V^{B_s h_{1s}}_0$
& $0.51^{-0.01+0.02}_{+0.01-0.02}$
& $1.60^{-0.03-0.01}_{+0.03+0.01}$
& $1.05^{-0.06-0.04}_{+0.07+0.04}$
\\
$V^{B_s h_{1s}}_1$
& $0.17^{+0.01+0.00}_{-0.01-0.00}$
& $1.16^{-0.05-0.02}_{+0.05+0.02}$
& $0.56^{-0.06-0.03}_{+0.07+0.03}$
& $V^{B_s h_{1s}}_2$
& $-0.10^{+0.01-0.00}_{-0.01+0.00}$
& $1.52^{-0.02-0.01}_{+0.02+0.01}$
& $0.95^{-0.03-0.03}_{+0.05+0.03}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
}
\subsection{$ B(0^-)\to D^{1/2} , D^{3/2}$ Form Factors}
\begin{itemize}
\item From Tables~\ref{tab:B2D}, we notice that most of the
form factors are small and lie between 0.1 to 0.25, except
$V_0(0)$ and $V_1(0)$ for the transitions emitting $P^{3/2}_1$ states, for
which these lie between 0.5 to 0.6. In contrast with these, $B, B_s \to D, D_s$
form factors carry the highest values between 0.6 to 0.8.
\item Slope parameters carry positive values except $a$ for $V_1$ form factor. However,
these parameters controlling $q^2$ behavior for $V_2$ form factor for
transitions emitting $P^{3/2}_1$ states remains difficult to control in spite of
choosing the $q^2$ dependence given in Eq. (\ref{eq:FFpara}).
\item Reverse changes occur in the form factors due to the variation in $\beta$
values for initial and final mesons. Increase in $\beta$ for initial (final)
state meson tend to decrease (increase) the magnitude of the form factors,
and
\item Minor changes occur in $B \to D$ form factors
from their previous values given in the earlier work \cite{Cheng04}, however,
slope parameters show significant difference.
\item To determine the physical form factors for $ B \to D_1$ transitions, one may need
the mixing angle between $D^{1/2}_1$ and $D^{3/2}_1$ states. A mixing angle
$ \theta_{D_1} = (5.76 \pm 2.4)^\circ$ is obtained by Belle through a detailed
$B \to D^* \pi \pi$ analysis \cite{Abe04}, while $ \theta_{D_{s1}} \approx 7^\circ$
is determined from the quark potential model \cite{Cheng03}.
\end {itemize}
\begin{table}[b]
\caption{Form factors of
$ B\rightarrow D^{1/2}_1, D^{3/2}_1$ transitions
obtained in the covariant light-front model are fitted to the
3-parameter form Eq. (\ref{eq:FFpara}). All the form factors are dimensionless.}
\label{tab:B2D}
\begin{ruledtabular}
\begin{tabular}{| c c c c |}
$F$
& $F(0)$
& $a$
& $b$
\\
\hline
$A^{B D^{1/2}_1}$
& $-0.13^{+0.01-0.02}_{-0.01+0.02}$
& $0.85^{-0.09+0.10}_{+0.08-0.18}$
& $0.12^{-0.01-0.01}_{+0.02+0.03}$
\\
$V^{B D^{1/2}_1}_0$
& $0.11^{-0.01+0.03}_{+0.01-0.03}$
& $1.08^{-0.02-0.02}_{+0.02-0.07}$
& $0.08^{-0.03+0.02}_{+0.03-0.04}$
\\
$V^{B D^{1/2}_1}_1$
& $-0.19^{+0.02-0.01}_{-0.02+0.01}$
& $-1.37^{-0.09-0.01}_{+0.08-0.00}$
& $1.07^{+0.07+0.01}_{-0.06-0.01}$
\\
$V^{B D^{1/2}_1}_2$
& $-0.14^{+0.02-0.02}_{-0.02+0.02}$
& $0.84^{-0.11+0.13}_{+0.09-0.21}$
& $0.13^{-0.01-0.01}_{+0.02+0.04}$
\\
\hline
$A^{B D^{3/2}_1}$
& $0.25^{-0.01+0.02}_{+0.01-0.02}$
& $1.17^{-0.03+0.01}_{+0.03-0.03}$
& $0.33^{-0.02-0.01}_{+0.02+0.01}$
\\
$V^{B D^{3/2}_1}_0$
& $0.52^{-0.01+0.04}_{+0.01-0.05}$
& $1.14^{-0.04+0.02}_{+0.03-0.06}$
& $0.34^{-0.02-0.01}_{+0.02+0.01}$
\\
$V^{B D^{3/2}_1}_1$
& $0.58^{-0.01+0.02}_{+0.01-0.03}$
& $-0.25^{-0.01-0.03}_{+0.01+0.02}$
& $0.29^{-0.00+0.01}_{+0.01-0.01}$
\\
$V^{B D^{3/2}_1}_2$
& $-0.10^{+0.01-0.02}_{-0.01+0.04}$
& $-5.95^{-2.07+3.80}_{+1.45-14.67}$
& $26.2^{+5.8-4.1}_{-11.5+41.0}$
\\
\hline
$A^{B_s D^{1/2}_{s1}}$
& $-0.17^{+0.02-0.02}_{-0.02+0.02}$
& $0.97^{-0.10+0.06}_{+0.10-0.10}$
& $0.37^{-0.05-0.04}_{+0.06+0.05}$
\\
$V^{B_s D^{1/2}_{s1}}_0$
& $0.13^{-0.01+0.03}_{+0.02-0.03}$
& $1.14^{-0.04-0.02}_{+0.04-0.05}$
& $0.29^{-0.04-0.03}_{+0.05+0.04}$
\\
$V^{B_s D^{1/2}_{s1}}_1$
& $-0.25^{+0.03-0.01}_{-0.03+0.01}$
& $-1.20^{-0.10-0.06}_{+0.10+0.07}$
& $1.02^{+0.07+0.04}_{-0.06-0.05}$
\\
$V^{B_s D^{1/2}_{s1}}_2$
& $-0.17^{+0.02-0.02}_{-0.02+0.02}$
& $0.96^{-0.12+0.08}_{+0.11-0.12}$
& $0.39^{-0.04-0.04}_{+0.06+0.06}$
\\
\hline
$A^{B_s D^{3/2}_{s1}}$
& $0.24^{-0.01+0.02}_{+0.01-0.02}$
& $1.26^{-0.06+0.00}_{+0.06-0.02}$
& $0.60^{-0.06-0.06}_{+0.07+0.06}$
\\
$V^{B_s D^{3/2}_{s1}}_0$
& $0.49^{-0.02+0.04}_{+0.02-0.05}$
& $1.25^{-0.06+0.01}_{+0.05-0.04}$
& $0.63^{-0.05-0.06}_{+0.06+0.07}$
\\
$V^{B_s D^{3/2}_{s1}}_1$
& $0.57^{-0.01+0.03}_{+0.01-0.04}$
& $-0.11^{-0.02-0.04}_{+0.02+0.04}$
& $0.32^{-0.01+0.00}_{+0.01+0.01}$
\\
$V^{B D^{3/2}_{s1}}_2$
& $-0.09^{+0.01-0.02}_{-0.01+0.03}$
& $-4.08^{-1.86+2.63}_{+1.24-7.55}$
& $21.1^{+5.3-8.3}_{-3.6+21.8}$
\\
\end{tabular}
\end{ruledtabular}
\end{table}
\section{Summary and conclusions}
In this work, we have studied the decay constants and form factors
of the ground-state $s$-wave and low-lying $p$-wave mesons within
a covariant light-front (CLF) approach. In the previous work \cite{Cheng04}, main
ingredients of the CLF quark model were explicitly worked out for both
$s$-wave and $p$-wave mesons. Besides that various form factors of the $D$ and $B$ mesons,
appearing in their transitions to isovector and isospinor $s$-wave and $p$-wave mesons,
were calculated within the framework of the CLF model.
In the present work, we have updated our results for these mesons, and
extended the analysis to determine the form factors for
$D_s$ and $B_s$ transitions, and also include the flavor-diagonal isoscalar
final states. Calculating the decay constants of most of the $s$-wave mesons
and a few axial vector mesons from the available experimental
data for various weak or electromagnetic decays, we
have fixed the shape parameter $\beta$ of the respective mesons, which
in turn determine the form factors. A few lattice results are
also used for this purpose. Errors in the $\beta$ parameters are
fixed from the corresponding experimental errors, otherwise standard
$10 \%$ uncertainty is assigned to investigate the effects of variation
in the $\beta$ parameter.
We have then proceeded to obtain the form factors in the CLF quark model for heavy-to-heavy and
heavy-to-light transitions of the charmed and bottom mesons to the
pseudoscalar mesons, vector mesons, scalar mesons and axial vector mesons. The $q^2$ dependence
of the form factors, generally assumed to be
given by Eq. (\ref{eq:FFpara}), is expressed through the slope
parameters, $a$ and $b$. Their sensitivity to the errors and the assigned
uncertainties of the $\beta$ parameters is investigated separately for the initial and the
final mesons.
Our main results are as follows:
\begin{itemize}
\item For $P \to P$ transitions, $B_s$
form factors at $q^2= 0$ are similar to that of the $B$ meson, as if
the spectator quark does not seem to affect them. Particularly, we observe
$F^{B_s D_s} = F^{B D}$, $F^{B_s K}\approx F^{B \pi}$, and
$F^{B_s \eta_s}\approx F^{B \eta_q}$, where
$\eta_{q}= (u\overline{u} + d\overline{d})/\surd {2}$, and
$\eta_{s}$ is pure $(s\overline{s})$ state. To lesser extant, the charmed mesons also show
a similar trend through $F^{D_s K} = F^{D \pi}$ and $F^{D_s \eta_s}\approx F^{D K}$. Heavy-to-light
form factors of the bottom mesons are smaller (around 0.3) than
that of the charmed mesons, which are around 0.7. The form factor
$F^{PP}_0$ generally shows a monopole behavior, and $F^{PP}_1$
acquires a dipole behavior.
\item For $P \to V$ transitions also, we find
$F^{B_s D^*_s}\approx F^{B D^*}$, $F^{B_s \phi}\approx F^{B \rho}\approx F^{B \omega}$,
$F^{D_s \phi}\approx F^{D K^*}$ and $F^{D_s K^*}\approx F^{D \rho}\approx F^{D \omega}$,
where $F$ denotes any of the four form factors, $V, A_0, A_1$ and $A_2$, at $q^2= 0$.
For the bottom mesons, heavy-to-light form factors are
smaller (from 0.2 to 0.4) than their heavy-to-heavy ones, which lie between 0.6 to 1.
Due to the reliability in fixing the $\beta$ parameters for the $s$-wave
mesons, the form factors at $q^2= 0$ hardly show sensitivity
to the errors in the $\beta$ values, though slope parameters ($a$ and $b$) generally
tend to increase (decrease) with decrease (increase) in $\beta$ for
the initial meson as well as the final meson.
\item Comparing $ P \to P, V$ form factors obtained
here with the results of other works, BSW model \cite{BSW}, the
Melikhov-Stech (MS) model \cite{Melikhov}, QCD sum rule
(QSR) \cite{Ball91}, light-cone sum rules (LCSR) \cite{LCSR},
lattice calculations \cite{LatticeFF} and perturbative
QCD approach \cite{RHLi09}, it is found that our form
factors agree well with the available lattice results, and
are most close to that of the MS model, except for the $B_s$
transitions. The LCSR and BSW model results are usually larger
for $P \to V$ form factors for $D$ and $B$ transitions, whereas
the QSR and pQCD calculations are generally lower than our results.
\item For $P \to S$ transitions, we have calculated the form factors
involving heavy scalar mesons only. These form factors, though are smaller
than the corresponding $P \to P$ form factors, also satisfy
$F^{B_s K^*_0} = F^{B a_0} = F^{B f_{0q}}$,
$F^{B_s f_{0s}} \approx F^{B D^*_0} = F^{B K^*_0}$, and
$F^{D_s f_{0s}}\approx F^{D a_0} = F^{D f_{0q}}$.
All the bottom meson form factors, lying between 0.25 and 0.30,
are roughly half of that of the charmed mesons, which are
around 0.5-0.6. The suppression of the $D_0^{*0}\pi^-$
production relative to $D^0\pi^-$ one clearly favors a smaller
$B\to D_0^*$ form factor relative to the $B\to D$ one.
The form factor $F^{PS}_0$ shows a monopole behavior, and $F^{PS}_1$
has a dipole behavior in general. The $P \to S$ form factors
are found to increase slowly with $q^2$ compared to the $P \to P $
ones. For the bottom sector,
the slope parameters are larger in magnitude than that for the charm sector.
These parameters (except for the case of negative $a$)
show an increase (decrease) with decrease (increase) in $\beta$ for each
of the initial and final mesons.
\item For heavy-to-light $P \to A(1^{++})$ transitions, the form
factor $A^{PA}(0)$ for the bottom mesons, generally lying around 0.25,
is larger than that for charmed meson transitions for which it
lies close to 0.16. Similarly, the form factor $V^{PA}_2(0)$ is $<0.1$
for the charm sector, where as it lies around 0.2 for the
bottom transitions. In contrast, the form factor $V^{PA}_1(0)$,
lying around 0.4 for the bottom sector is significantly
smaller than that for the charm sector, where its value
lies between 1.4 and 1.8. Also $V^{PA}_0(0)$ for the bottom
transitions is roughly half of its value for the charmed
meson transitions. We also observe that
$V^{PA}_1 > V^{PA}_0 > A^{PA} > V^{PA}_2$ for the charmed mesons
and $V^{PA}_1 > A^{PA} > V^{PA}_2 > V^{PA}_0$ for the bottom mesons.
All the slope parameters, except
for $V^{PA}_1$ and $V^{PA}_2$ for the charmed meson
transitions, are found to be positive, and for the
bottom mesons, their values are significantly
larger than that for the charm sector.
\item For heavy-to-light transitions $P \to A(1^{+-}) $,
the form factors, $A^{PA}(0), V^{PA}_0(0)$ and $V^{PA}_1(0)$, are positive,
where as $V^{PA}_2(0)$ is negative and small (around -0.1) for
the bottom as well as the charmed sector. These follow the pattern:
$V^{PA}_1 > V^{PA}_0 > A^{PA} > |V^{PA}_2| $ for the charmed mesons
and $V^{PA}_0 > V^{PA}_1 > A^{PA} \geq |V^{PA}_2| $ for the bottom mesons.
Numerically, the form factor
$A^{PA}(0)$ is generally around 0.1, where as $V^{PA}_0(0)$ lies close
to 0.5 for all the cases. Form factor $V^{PA}_1(0)$, lying between
0.15 to 0.20, for the bottom sector is significantly smaller than
that for the charm sector, where its value lies between
1.3 to 1.6. Typically for the heavy-to-light $P \to A(1^{+-})$ transitions,
the form factors $A^{PA}(0)$ and $V^{PA}_1(0)$ usually increase (decrease) with
increasing (decreasing) beta for initial meson as well as for
final meson. Both $V^{PA}_0(0)$ and $V^{PA}_2(0)$ form factors
decrease (increase) in magnitude
with increasing (decreasing) $\beta$ for the initial mesons, and show the opposite
trend for the final mesons. For the bottom sector,
the slope parameters are found to be larger than that for the charm sector.
\item For $B \to D^{1/2}_1/D^{3/2}_1$ transitions, all the
form factors lie between 0.1 to 0.2, except for
$V_0(0)$ and $V_1(0)$ for the transitions emitting $P^{3/2}_1$
states, for which these lie between 0.5 to 0.6. Slope parameters
carry positive values, except
$a$ for the $V_1$ form factor. Reverse changes occur in the
form factors due to variation in $\beta$ values for
initial and final mesons, i.e., increase in $\beta$ for the initial (final)
state meson tend to decrease (increase) the magnitude of the form factors.
\item Now several model calculations for
$B\rightarrow A$ form factors are available: the ISGW2 model \cite{ISGW2},
the constituent quark-meson model (CQM) \cite{Deandrea},
the QCD sum rules (QSR) \cite{Aliev}, light cone sum rules
(LCSR) \cite{KCYang08}, and the
perturbative QCD (pQCD) approach \cite{RHLi09}.
Significant differences are observed, since these approaches differ in
their treatment of dynamics of the form factors.
For instance, $V^{Ba_1}_0 = 1.20$, obtained in the CQM model, and $1.01$ in the ISGW2 model,
is much larger than its values obtained in other approaches. The BaBar
and Belle measurements \cite{BABAR, Abe07} of
$\overline{B}^0\rightarrow a^{\pm}_1 \pi^{\mp}$ seem to favor a
value of $V^{Ba_1}_0\approx 0.30$ \cite{ChengAP}. We have earlier pointed out
\cite{Cheng04} that relativistic effects could manifest
in heavy-to-light transitions at maximum recoil where the final-state meson can be
highly relativistic, which can naturally be considered in the CLF model.
Various form factors, calculated using the CLF model,
have earlier been used to study weak hadronic and radiative decays
of the bottom mesons emitting $p$-wave mesons \cite{Cheng07, Cheng10},
and a good agreement between theory and available experimental
data could be obtained. It has been pointed out in the previous work \cite{Cheng04} that
the requirement of HQS is also satisfied for
the decay constants and the form factors obtained in the CLF quark model. Particularly, it has been
shown that the Bjorken \cite{Bjorken} and Uraltsev \cite{Uraltsev} sum rules for the
Isgur-Wise functions are satisfied.
\end{itemize}
\vskip 2.5cm \acknowledgments RCV thanks
the Institute of Physics, Academia Sinica and National Center for
Theoretical Sciences, National Tsing-Hua University, for their hospitality
during his visits, where most part of this work was done. He specially acknowledges
H.Y. Cheng for the invitation, useful discussions and reading the manuscript.
He also expresses his thanks to C.K. Chua for providing his source codes. This research
was supported in part by the National Science Council of R.O.C. under
No. NSC97-2112-M-001-004-MY3 and by the special NCTS grant of Academia Sinica.
\eject
|
\section{Introduction}\label{sec:1}
The theory of backward stochastic differential equations (BSDEs) and
that of fractional Brownian motion (fBm) had developed
simultaneously in their own separated directions for many years
until Bender \cite{Be} gave an explicit solution to a linear BSDE
driven by fBm in 2005. In 2009 Hu and Peng \cite{HP} stated a more
general theory on fractional BSDEs by using the so-called
quasi-expectation, but their case is still limited. The
non-semimartingale property of the fBm (except the case of Hurst
parameter $H=1/2$, where it becomes a Brownian motion) makes it
thorny to handle. Being not a semimartingale means there is no
martingale representation theory for the fBm, which is crucial in
the general BSDE theory (see the pioneering work of Pardoux and Peng
\cite{PP3}). In Jing and Le\'on \cite{Jl}, we tried to combine the
fBm and BSDEs in another way: We transformed a semilinear backward
doubly stochastic differential equation (BDSDE) driven by both a
standard and a fractional Brownian motions with $H\in (0,1/2)$ into
a BSDE without integral with respect to the fBm, which turns out
easier to deal with. The integral w.r.t. the fBm in the BDSDE was
interpreted in the sense of the extended divergence operator and the
integral was supposed to be linear w.r.t. the solution process. This
allowed to apply the efficient tool of nonanticipating Girsanov
transformation, developed in Buckdahn \cite{bu}. However, this
method is restricted to semilinear BDSDEs.
In this paper, we deal with BDSDEs for which the integrand of the
integral with respect to the fBm is not necessarily linear with the
solution process, and the Hurst coefficient $H$ is supposed to
belong t o the interval$(1/2,1)$. Unlike the more irregular case
$H<1/2$, the stochastic integrals with respect to an fBm with
$H>1/2$ can be defined in different ways. So they can be defined
with the help of the divergence operator in the frame of the
Malliavin calculus, see Decreusefont and \"Ust\"unel \cite{Du} and
Al\`{o}s {\it et al.} \cite{Amn} (Notice that the Wick-It\^o
integral defined in Duncan {\it et al.} \cite{Dhp} coincides with
the first one). They can also be defined pathwise as generalized
Riemann-Stieltjes integral (see Z\"ahle \cite{Za1} and \cite{Za2})
or with the help of the rough path theory (see Coutin and Qian
\cite{Cq}). For a complete list of references we refer to the two
books by Biagini {\it et al.} \cite{BHOZ} and Mishura \cite{Mi}.
Our approach to BDSDEs with an fBM is inspired by the work of
Buckdahn and Ma \cite{BM}. In their study of stochastic PDEs driven
by a Brownian motion $B$ the authors of \cite{BM} used BDSDEs
driven by $B$ as well as an independent Brownian motion; the
integral with respect to $B$ is interpreted in Stratonovich sense.
This allowed the application of the Doss-Sussman transformation in
order to transform the BDSDE into a BSDE without integral with
respect to $B$. On the other hand, the pathwise integral with
respect to the fBm plays a role which is comparable with that of the
Stratonovich integral in the classical theory. Nualart and
R\u{a}\c{s}canu \cite{Nr} used the pathwise integral to solve
(forward) stochastic differential equations driven by an fBm. For
some technical reasons (such as the lack of H\"older continuity, see
Remark \ref{rmk2}), we shall make use of the Russo-Vallois integral
developed by Russo and Vallois in a series of papers (\cite{Rv1},
\cite{Rv1}, \cite{Rv1}, {\it etc}.). Under standard assumptions
which allow to apply the Doss-Sussman transformation, we associate
the BDSDE driven by both a standard Brownian motion $W$ and an fBm
$B$,
\begin{equation}\label{eq:1.2}
U_s^{t,x}=\Phi(X_0^{t,x})+\int^s_0f(r,X_r^{t,x},
U_r^{t,x},V_r^{t,x})\d r + \int^s_0 g(U_r^{t,x})\d
B_r-\int^s_0V_r^{t,x}\dw\d W_r, \ s\in[0,t], \footnote{$\int^t_0
\cdot\dw \d W_s$ \textrm{indicates that the integral is considered
as the It\^o backward one.}}
\end{equation}
with the BSDE driven only by the Brownian motion $W$,
\begin{equation} \label{eq:1.1}
Y_s^{t,x}=\Phi(X_0^{t,x})+\int^s_0\tilde{f}(r,X_r^{t,x},Y_r^{t,x},Z_r^{t,x})\d
r -\int^s_0 Z_r^{t,x}\dw \d W_r, \ s\in[0,t].
\end{equation}
Here $\tilde{f}$ will be specified in Section 4; it is a driver
with quadratic growth in $z$. We point out that the classical BDSDEs
were first studied by Pardoux and Peng \cite{PP2} and our BSDE
(\ref{eq:1.1}) is a quadratic growth BSDE, which was studied first
by Kobylanski \cite{Ko}. In the works of Pardoux and Peng \cite{PP2}
and Buckdahn and Ma \cite{BM}, i.e., when the Hurst parameter
$H=1/2$, one can solve the BDSDE directly and get the square
integrability of the solution process. However, in the fractional
case $(H\neq1/2)$, to our best knowledge, there does not exist a
direct way to solve the BDSDE (\ref{eq:1.2}), and as it turns out in
Theorem \ref{thm:sol}, we can only get that the conditional
expectation of $\int^t_0\lt|Z^{t,x}_r\rt|^2\d r$ is bounded by an
a.s. finite process. This is also the reason that instead of using
the space of square integrable processes, we use the space of a.s.
conditionally square integrable processes (see the definition of the
space $\HH_T^2(\R^d)$ in Section 2).
A celebrated contribution of the BSDE theory consists in giving a
form of probabilistic interpretation, nonlinear Feynman-Kac formula,
to the solutions of PDEs (see, for instance, Peng \cite{Pe}, Pardoux
and Peng \cite{PP1}). As indicated in Kobylanski \cite{Ko}, the
quadratic growth BSDE (\ref{eq:1.1}) is connected to the semilinear
parabolic PDE
\begin{equation}\label{eq:1.3}
\left\{
\begin{array}{cllc}
\frac{\partial u}{\partial t}(t,x)&=&\mathscr{L} u (t,x) -
\tilde{f}(t,x,u(t,x),\sigma(x)^T\frac{\partial}{\partial x}u(t,x)),&
(t,x)\in (0,T)\times
\R^n;\\
u(0,x)&=&\Phi(x),&x\in\R^n,
\end{array}\right.
\end{equation}
where $\mathscr{L}$ is the infinitesimal operator of a Markov
process. Hence, it is natural for us to consider the form of equation
(\ref{eq:1.3}) after Doss-Sussman transformation and we prove that
it becomes the following semilinear SPDE
\begin{equation}\label{eq:1.4}
\left\{
\begin{array}{lc}
\d u(t,x)=\left[\mathscr{L} u (t,x) -
{f}(t,x,u(t,x),\sigma(x)^T\frac{\partial}{\partial x}u(t,x))\rt]\d t
+ g(u(t,x))\d B_t,& (t,x)\in (0,T)\times
\R^n;\\
u(0,x)=\Phi(x),&x\in\R^n.
\end{array}\right.
\end{equation}
We emphasize that this paper can not be considered as a
generalization of \cite{Jl}. The reason is that, firstly, the Hurst
parameters are distinct; secondly, the stochastic integrals with
respect to the fBm are of different types.
We organize the paper as follows. In section 2, we recall some basic
facts about the fractional Brownian motion, we give the general
framework of our work, and we recall the definition of the backward
Russo-Vallois integral as well as some of its properties. In section
3 we prove a type of It\^o formula involving integrals with respect
to both standard and fractional Brownian motions, which will play an
important role in the following sections. We perform a Doss-Sussman
transformation in Section 4 to transform a nonlinear BDSDE
(\ref{eq:1.2}) into a BSDE (\ref{eq:1.1}) and show the relationship
between their solutions. In particular, we show that BSDE
(\ref{eq:1.1}) has a unique solution $(Y^{t,x},Z^{t,x})$, and the
couple of processes $(U^{t,x},V^{t,x})$ associated with
$(Y^{t,x},Z^{t,x})$ by the inverse Doss-Sussman transformation is
the unique solution of BDSDE ({\ref{eq:1.2}}). Finally, the
stochastic PDE associated with BDSDE (\ref{eq:1.2}) is briefly
discussed in Section 5.
\section{Preliminaries}
\subsection{Fractional Brownian Motion and General Setting}
In this subsection we recall some basic results on the fBm and the
related setting. For a more complete overview of the theory of fBm,
we refer the reader to Biagini {\it et al.} \cite{BHOZ} and Mishura
\cite{Mi}.
Let $(\Omega^\pr,\F^\pr, \PP^\pr)$ be a classical Wiener space with
time horizon $T>0$, i.e., $\Omega^\pr=C_0([0,T];\R)$ denotes the set
of real-valued continuous functions starting from zero at time zero,
endowed with the topology of the uniform convergence,
$\mathcal{B}(\Omega^\pr)$ is the Borel $\sigma$-algebra on
$\Omega^\pr$ and $\PP^\pr$ is the unique probability measure on
$(\Omega, \mathcal{B}(\Omega^\pr))$ with respect to which the
coordinate process $W^0_t(\omega^\pr)=\omega^\pr(t),$ $t\in[0,T],$
$\omega^\pr\in\Omega^\pr$ is a standard Brownian motion. By $\F^\pr$
we denote the completion of $\mathcal{B}(\Omega^\pr)$ by all
$\PP^\pr$-null sets in $\Omega^\pr$. Given $H\in(1/2,1)$, we define
$$
B_t=\int_0^t K_H(t,s)\d W^0_s,\quad t\in[0,T],
$$
where $K_H$ is the kernel of the fBm with parameter $H\in(1/2,1)$:
$$
K_H(t,s)=C_Hs^{1/2 - H}\int_s^t u^{H-1/2}(u-s)^{H-3/2}\d u,
$$
with $C_H =\sqrt{\frac{H}{(2H-1)\beta (2-2H, H - {1}/{2})}}$. It is
well known that such defined process $B$ is a one-dimensional fBm,
i.e., it is a Gaussian process with zero mean and covariance
function
$$
R_H(t,s):= \mathbb{E}\left[B_tB_s\right] = \frac{1}{2}\left(t^{2H}+
s^{2H}-|t-s|^{2H}\right), \quad s, t\in [0,T].
$$
We let $\{W_t: {0\leq t\leq T}\}$ be the coordinate process on the
classical Wiener space $(\Omega^{\pr\pr},{\cal F} ^{\pr\pr},
\mathbb{P}^{\pr\pr})$ with $\Omega^{\pr\pr}=C_0([0,T]; \R^d)$, which
is a $d$-dimensional Brownian motion with respect to the Wiener
measure $\PP^{\pr\pr}$. We put $(\Omega, {\cal F}^0,
\mathbb{P})=(\Omega^{\pr},{\cal F}^{\pr}, \mathbb{P}^{\pr})
\otimes(\Omega^{\pr\pr},{\cal F} ^{\pr\pr}, \mathbb{P}^{\pr\pr})$
and let $\mathcal {F}=\mathcal{F}^0\vee \mathcal{N}$, where
$\mathcal N$ is the class of the $ \mathbb{P}$-null sets. We denote
again by $B$ and $W$ the canonical extensions of the fBm $B$ and of
the Brownian motion $W$ from $\Omega^{\pr}$ and $\Omega^{\pr\pr}$,
respectively, to $\Omega$.
We let $\mathcal{F}_{[t,T]}^W = \sigma\{W_T-W_s, t\leq s\leq
T\}\vee\cal N$, $\mathcal{F}_t^{B}=\sigma\{B_s, 0\leq s\leq
t\}\vee\cal N,$ and $\mathcal{G}_t = \mathcal{F}_{[t,T]}^W
\vee\mathcal{F}_t^{B}, t\in[0,T].$ Let us point out that
$\mathcal{F}_{[t,T]}^W$ is decreasing and $\mathcal{F}_t^{B}$ is
increasing in $t$, but $\mathcal{G}_t$ is neither decreasing nor
increasing. We denote the family of $\sigma$-fields
$\{\mathcal{G}_t\}_{0\leq t\leq T}$ by $\mathbb{G}$. Moreover, we
also introduce the backward filtrations $\mathbb{H}=\{\mathcal {H}_t
=\mathcal{F}_{[t,T]}^W \vee \mathcal{F}_T^{B}\}_ {t\in[0,T]}$ and
$\mathbb{F}^W=\{\mathcal{F}_{[t,T]}^W\}_{t\in[0,T]}$.
Finally, we denote by $C(\mathbb{H},[0,T]; \R^m)$ the space of the
$\R^m$-valued continuous processes $\{\varphi_t, t\in[0,T]\}$ such
that $\varphi_t$ is $\mathcal{H}_t$-measurable, $t\in[0,T]$, and
$\mathcal{M}^2(\mathbb{F}^W,[0,T];\R^m)$ the space of the
$\R^m$-valued square-integrable processes $\{\psi_t, t\in[0,T]\}$
such that $\psi_t$ is $\mathcal{F}_{[t,T]}^W$-measurable,
$t\in[0,T]$. Let $\HH_T^{\infty}(\R)$ be the set of
$\mathbb{H}$-progressively measurable processes which are almost
surely bounded by some real-valued $\F_T^B$-measurable random
variable, and let $\HH_T^2(\R^d)$ denote the set of all
$\R^d$-valued $\mathbb{H}$-progressively measurable processes
$\gamma=\{\gamma_t: t\in[0,T]\}$ such that
$\E\lt[\int^{T}_0|\gamma_t|^2\d t|\F^B_{T}\rt]<+\infty$, $\PP$-a.s.
\subsection{Russo-Vallois Integral}\label{sec:RV}
\setcounter{equation}{0} In a series of papers (\cite{Rv},
\cite{Rv1}, \cite{Rv2}, {\it etc.}), Russo and Vallois defined new
types of stochastic integrals, namely forward, backward and
symmetric integrals, which are extensions of the classical
Riemann-Stieltjes integral, and in fact these three integrals
coincide, when the integrator is a fBm with Hurst parameter
$H\in(1/2, 1)$. Here we will mainly use the backward Russo-Vallois
integral in this paper. It turns out to be a convenient definition
for stochastic integral with respect to our fBm $B$.
Let us recall some results by Russo and Vallois which we will use
later. In what follows, we make the convention that all continuous
processes $\{X_t, t\in[0,T]\}$ are extended to the whole line by
putting $X_t=X_0$, for $t<0$, and $X_t=X_T$, for $t>T$.
\begin{definition}{\label{def:rv}}
Let $X$ and $Y$ be two continuous processes. For $\ep>0$, we set
$$
I(\ep,t,X,\d Y)\triangleq\frac{1}{\ep}\int^t_0X(s)(Y(s)-Y(s-\ep))\d s,
$$
\[
C_\ep(X,Y)(t)\triangleq\frac{1}{\ep}\int^t_0(X(s)-X(s-\ep))(Y(s)-Y(s-\ep))\d
s, \ t\in[0,T].
\]
Then the backward Russo-Vallois integral is defined as the uniform
limit in probability as $\ep\to0^+$, if the limit exists. The
generalized bracket $[X,Y]$ is the uniform limit in probability of
$C_{\ep}(X,Y)$ as $\ep\to0^+$ $($of course, again under the
condition of existence$)$.
\end{definition}
We recall that (cf. Protter \cite{Pr}) a sequence of processes
$(H_n;n\ge0)$ converges to a process $H$ uniformly in probability if
\[
\lim_{n\to\infty}\mathbb{P}\lt(\sup_{t\in[0,T]}|H_n(t)-H(t)|>\alpha\rt)=0
\quad \mbox{for\ \ every}\ \ \alpha>0.
\]
In \cite{Rv2} (Theorem 2.1) Russo and Vallois derived the It\^o
formula for the backward Russo-Vallois integral.
\begin{theorem}\label{thm:russo}
Let $f\in C^2(\R)$ and $X$ be a continuous process admitting the
generalized bracket, i.e., $[X,X]$ exists in the sense of Definition
\ref{def:rv}. Then for every $t\in[0,T]$, the backward Russo-Vallois
integral $\int^t_0f^\pr(X(s))\d X(s)$ exists and
\[
f(X(t))=f(X(0))+\int^t_0f^\pr(X(s))\d X(s)-\frac{1}{2}\int^t_0F^{\pr\pr}
(X(s))\d [X,X](s),
\]
for every $t\geq0$.
\end{theorem}
We list some properties of Russo-Vallois integral, which will be
used later in this paper.
\begin{proposition}\label{prop_rv1}
$(1)$. If $X$ is a finite quadratic variation process $($i.e.,
$[X,X]$ exists and $[X,X]_T<+\infty$, $\PP$-a.s.$)$ and $Y$ is a
zero quadratic variation process $($i.e., $[Y,Y]$ exists and equals
to zero$)$, then the mutual generalized bracket $[X,Y]$ exists and
vanishes, $\PP$-a.s.
$(2)$. If $X$ and $Y$ have $\PP$-a.s. H\"older continuous paths with
order $\alpha$ and $\beta$, respectively, such that $\alpha>0,$
$\beta>0$ and $\alpha+\beta>1$, then $[X,Y]=0$.
$(3)$. We assume that $X$ and $Y$ are continuous and admit a mutual
bracket. Then, for every continuous process $\{H(s):s\in[0,T]\}$,
$$
\int^\cdot_0H(s)\d C_\ep(X,Y)(s)\ \ \mbox{converges\ to}\ \
\int^\cdot_0H(s)\d [X,Y](s).
$$
\end{proposition}
The following proposition, which can be found in Russo and Vallois
\cite{Rv}, states the relationship between the Young integral (see
Young \cite{Yo}) and the backward Russo-Vallois integral.
\begin{proposition}
Let $X, Y$ be two real processes with paths being $\PP$-a.s. in
$C^\alpha$ and $C^\beta$, respectively, with $\alpha>0$, $\beta>0$
and $\alpha+\beta>1$. Then the backward Russo-Vallois integral
$\int^\cdot_0 Y\d X$ coincides with the Young integral
$\int^{\cdot}_0 Y\d^{(y)}X$.
\end{proposition}
\section{A Generalized It\^o Formula}
\setcounter{equation}{0} In this section we state a generalized
It\^o formula involving an It\^o backward integral with respect to
the Brownian motion $W$ and the Russo-Vallois integral with respect
to the fBm $B$. It will play an important role in our paper. It is
noteworthy that this It\^o formula corresponds to Lemma 1.3 in the
paper of Pardoux and Peng \cite{PP2} for the case of an fBm with
Hurst parameter $H=1/2$, i.e., when $B$ is a Brownian motion.
\begin{theorem}\label{thm:ito}
Let $\alpha\in C(\mathbb{H},[0,T]; \R)$ be a process of the form
\[
\alpha_t=\alpha_0+\int^t_0\beta_s\d s +\int^t_0 \gamma_s\dw\d W_s,
\quad t\in[0,T],
\]
where $\beta$ and $\gamma$ are $\mathbb{H}$-adapted processes and
$\PP\{\int^T_0|\beta_s|\d s<\infty\}=1$ and
$\PP\{\int^T_0|\gamma_s|^2\d s<+\infty\}=1$, respectively. Suppose
that $F\in C^2(\R\times\R)$. Then the Russo-Vallois integral
$\int_0^t\frac{\partial F}{\partial y}(\alpha_s,B_s)\d B_s $
$($defined as the uniform limit in probability of
$\frac{1}{\ep}\int^t_0(B_s-B_{s-\ep})\frac{\partial F}{\partial
y}(\alpha_s,B_s)\d s$$)$ exists for $0\le t\le T$,
and it holds that, $\mathbb{P}$-almost surely, for all $0\le t\le
T$,
\begin{equation}\label{eq:ito}
\begin{aligned}
F(\alpha_t,B_t)=&F(\alpha_0,0)+\int^t_0 \frac{\partial F}{\partial
x}(\alpha_s,B_s)\beta_s\d s +\int^t_0 \frac{\partial F}{\partial
x}(\alpha_s,B_s)\gamma_s\dw\d W_s + \int^t_0 \frac{\partial
F}{\partial
y}(\alpha_s,B_s)\d B_s\\
&-\frac12\int^t_0 \frac{\partial^2 F}{\partial
x^2}(\alpha_s,B_s)|\gamma_s|^2\d s.
\end{aligned}
\end{equation}
\end{theorem}
\noindent\textit{Proof:} {\bf{Step 1}}. First we suppose $F\in
C^2_b(\R\times \R)$ (i.e., the function $F$ is twice continuously
differentiable and has bounded derivatives of order less than or
equal to two) and there is a positive constant $C$ such that
$\int^T_0|\beta_s|\d s\le C$ and $\int^T_0|\gamma_s|^2\d s\le C$. It
is direct to check that
$$
F(\alpha_t,B_t)-F(\alpha_0,0)=\lim_{\ep\to0}\frac{1}{\ep}\int^t_0
(F(\alpha_s,B_s)-F(\alpha_{s-\ep},B_{s-\ep}))\d s.
$$
For simplicity we put $\alpha_{a,\ep,s}\triangleq
\alpha_{s}-a(\alpha_s-\alpha_{s-\ep})$ and $B_{a,\ep,s}\triangleq
B_{s}-a(B_s-B_{s-\ep})$, for any $a\in[0,1]$, $s\in[0,T],$ $\ep>0$.
We have
\begin{equation}\label{eq:3.2}
\begin{aligned}
&F(\alpha_s,B_s)-F(\alpha_{s-\ep},B_{s-\ep})\\
= &(\alpha_s-\alpha_{s-\ep}) \frac{\partial F}{\partial
x}(\alpha_s,B_s)+(B_s-B_{s-\ep}) \frac{\partial F}{\partial
y}(\alpha_s,B_s)-(\alpha_s-\alpha_{s-\ep})^2\int^1_0
\frac{\partial^2 F}{\partial x^2} (\alpha_{a,\ep,s},
B_{a,\ep,s})(1-a)\d
a\\
&-2(\alpha_s-\alpha_{s-\ep})(B_s-B_{s-\ep})\int^1_0 \frac{\partial^2
F}{\partial x\partial y} (\alpha_{a,\ep,s}, B_{a,\ep,s})(1-a)\d
a\\
&-(B_s-B_{s-\ep})^2\int^1_0 \frac{\partial^2 F}{\partial y^2}
(\alpha_{a,\ep,s}, B_{a,\ep,s})(1-a)\d a.
\end{aligned}
\end{equation}
By applying the stochastic Fubini theorem, we get that
$$
\begin{aligned}
&\frac{1}{\ep}\int^t_0 (\alpha_s-\alpha_{s-\ep})
\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s\\
=&\frac{1}{\ep}\int^t_0 \lt(\int^s_{s-\ep}\beta_r\d r+\int^s_{s-\ep}
\gamma_r\dw\d W_r\rt)
\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s\\
=&\frac{1}{\ep}\int^t_0 \int_r^{(r+\ep)\wedge
t}\beta_r\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s\d r
+\frac{1}{\ep}\int^t_0 \int_r^{(r+\ep)\wedge t} \gamma_r
\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s\dw\d W_r.
\end{aligned}
$$
Since $\frac{1}{\ep}\int_r^{(r+\ep)\wedge t}\frac{\partial
F}{\partial x}(\alpha_s,B_s)\d s$ is $\HH_t$-measurable and
converges to $\frac{\partial F}{\partial x}(\alpha_r,B_r)$ when
$\ep\to0$, it follows that, thanks to the continuity of
$\frac{\partial F}{\partial x}(\alpha_s,B_s)$,
$$
\begin{aligned}
&\lim_{\ep\to0} \sup_{t\in[0,T]}\lt|\int^t_0 \beta_r \lt(
\frac{1}{\ep}\int_r^{(r+\ep)\wedge t}\frac{\partial F}{\partial
x}(\alpha_s,B_s)\d s-\frac{\partial F}{\partial
x}(\alpha_r,B_r)\rt)\d r\rt|\\
\le& \lim_{\ep\to0} \sup_{r\in[0,T]}\lt|
\frac{1}{\ep}\int_r^{(r+\ep)\wedge t}\frac{\partial F}{\partial
x}(\alpha_s,B_s)\d s-\frac{\partial F}{\partial x}(\alpha_r,B_r)\rt|
\int^T_0 |\beta_r|\d r= 0,\ \ \mbox{in\ \ probability}.
\end{aligned}
$$
Thus, in virtue of the boundedness of $\frac{\partial F}{\partial
x}$, by the Dominated Convergence Theorem,
$$
\begin{aligned}
&\lim_{\ep\to
0}\mathbb{E}\lt[\lt.\sup_{t\in[0,T]}\lt|\int^t_0\gamma_r\lt(\frac{1}{\ep}\int_r^{(r+\ep)\wedge
t} \frac{\partial F}{\partial x}(\alpha_s,B_s)\d s-\frac{\partial
F}{\partial x}(\alpha_r,B_r)\rt)\dw\d W_r\rt|^2\rt|\F_T^B\rt]
\\
\le & \lim_{\ep\to
0}\mathbb{E}\lt[\lt.\sup_{r\in[0,T]}\lt|\frac{1}{\ep}\int_r^{(r+\ep)\wedge
t} \frac{\partial F}{\partial x}(\alpha_s,B_s)\d s-\frac{\partial
F}{\partial x}(\alpha_r,B_r)\rt|^2\int^t_0|\gamma_r|^2\d
r\rt|\F_T^B\rt]= 0, \ \PP-\textrm{a.s.}
\end{aligned}
$$
(Recall that $\lt(\frac{1}{\ep}\int_r^{(r+\ep)\wedge t}
\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s-\frac{\partial
F}{\partial x}(\alpha_r,B_r)\rt)_{r\in[0,T]}$ is
$\mathbb{H}$-adapted and $W_\cdot-W_T$ is an $\mathbb{H}$-(backward)
Brownian motion.) Thus, we get
\begin{equation}\label{eq:3.3}
\begin{aligned}
\lim_{\ep\to0}\frac{1}{\ep}\int^t_0 (\alpha_s-\alpha_{s-\ep})
\frac{\partial F}{\partial x}(\alpha_s,B_s)\d s =\int^t_0\beta_r
\frac{\partial F}{\partial x}(\alpha_r,B_r)\d r +\int^t_0 \gamma_r
\frac{\partial F}{\partial x}(\alpha_r,B_r)\dw\d W_r,
\end{aligned}
\end{equation}
uniformly in probability. We notice that the generalized bracket of
$\alpha$ is the same as the classical one, i.e.,
$[\alpha,\alpha]_s=\int^s_0|\gamma_r|^2\d r$, $s\in[0,T]$. We also
have
\begin{equation}\label{eq:3.4}
\begin{aligned}
&\frac{1}{\ep}\int^t_0 (\alpha_s-\alpha_{s-\ep})^2\int^1_0
\frac{\partial^2 F}{\partial x^2} (\alpha_{a,\ep,s},
B_{a,\ep,s})(1-a)\d a \d s\\
=&\frac{1}{2\ep}\int^t_0 (\alpha_s-\alpha_{s-\ep})^2
\frac{\partial^2 F}{\partial x^2} (\alpha_s, B_s)\d s+ A_{\ep,t}, \
\ t\in[0,T],
\end{aligned}
\end{equation}
where
$$
A_{\ep,t}=\frac{1}{\ep}\int^t_0 (\alpha_s-\alpha_{s-\ep})^2 \int^1_0
\lt(\frac{\partial^2 F}{\partial x^2} (\alpha_{a,\ep,s},
B_{a,\ep,s})-\frac{\partial^2 F}{\partial x^2} (\alpha_s,
B_s)\rt)(1-a)\d a \d s.
$$
Proposition \ref{prop_rv1} yields that $ \frac{1}{2\ep}\int^._0
(\alpha_s-\alpha_{s-\ep})^2 \frac{\partial^2 F}{\partial x^2}
(\alpha_s, B_s)\d s $ converges to $\int^._0 \frac{\partial^2
F}{\partial x^2} (\alpha_s, B_s)\d [\alpha,\alpha]_s$ and the
continuity of $\frac{\partial^2 F}{\partial x^2}$, $\alpha$ and $B$
implies that $A_{\ep,t}$ converges to zero. A similar argument shows
that
$$
\frac{1}{\ep}\int^._0
2(\alpha_s-\alpha_{s-\ep})(B_s-B_{s-\ep})\int^1_0 \frac{\partial^2
F}{\partial x\partial y} (\alpha_{a,\ep,s}, B_{a,\ep,s})(1-a)\d a \d
s$$ and the term
$$
\frac{1}{\ep}\int^._0(B_s-B_{s-\ep})^2\int^1_0 \frac{\partial^2
F}{\partial y^2} (\alpha_{a,\ep,s}, B_{a,\ep,s})(1-a)\d a\d s
$$
converge in probability, respectively, to
$$
2\int^._0\frac{\partial^2 F}{\partial x\partial y} (\alpha_s, B_s)\d
[\alpha,B]_s\ \ \mbox{and} \ \ \int^._0\frac{\partial^2 F}{\partial
x\partial y} (\alpha_s, B_s)\d [B,B]_s.
$$
However, these both latter expressions are zero due to the fact that
$H\in(1/2,1)$ and Proposition \ref{prop_rv1} (Observe that the fBm
has H\"older continuous paths of any positive order less than $H$
almost surely).
Combining the above results with (\ref{eq:3.2}), (\ref{eq:3.3}) and
(\ref{eq:3.4}), we see that
$$
\begin{aligned}
\lim_{\ep\to0}\frac{1}{\ep}\int^t_0(B_s-B_{s-\ep})\frac{\partial
F}{\partial y}(\alpha_s,B_s)\d
s=&F(\alpha_t,B_t)-F(\alpha_0,0)-\int^t_0 \frac{\partial F}{\partial
x}(\alpha_s,B_s)\beta_s\d s \\&-\int^t_0 \frac{\partial F}{\partial
x}(\alpha_s,B_s)\gamma_s\dw\d W_s +\frac12\int^t_0 \frac{\partial^2
F}{\partial x^2}(\alpha_s,B_s)|\gamma_s|^2\d s,
\end{aligned}
$$
uniformly in $t\in[0,T]$, in probability. Consequently, the integral
$\int^t_0\frac{\partial F}{\partial y}(\alpha_s, B_s)\d B_s$ exists
in Russo-Vallois' sense (Recall Definition \ref{def:rv}) and we get
the It\^o formula (\ref{eq:ito}) for $F\in C^2_b(\R\times\R)$.
{\bf Step 2}. Now we deal with the general case that
$\PP\{\int^T_0|\beta_s|\d s<\infty \}=\PP\{\int^T_0|\gamma_s|^2\d
s<\infty\}=1$. For each $n\in\mathbb{N}$, we define a sequence of
$\mathbb{H}$-stopping times by $\tau_n=\sup\{t\le
T:\int^T_t|\beta_s|\d s+\int_t^T|\gamma_s|^2\d s>n\}\vee 0$, so we
know that the processes $\{\beta_t^n:=\beta_t
{\bf{1}}_{[\tau_n,T]}(t), t\in[0,T]\}$ and $\{\gamma^n_t:=\gamma_t
{\bf{1}}_{[\tau_n,T]}(t), t\in[0,T]\}$ satisfy
$\int^T_0|\beta^n_s|\d s\le n$ and $\int^T_0|\gamma^n_s|^2\d s\le
n,$ $\PP$-a.s. Furthermore, as $n\to\infty$, $\tau_n\to 0,$
$\PP$-a.s. We consider the It\^o formula for the process $
\alpha^n_t=\alpha_0+\int^t_0\beta^n_s\d s+ \int^t_0\gamma_s^n\dw \d
W_s, t\in[0,T]. $ Thanks to the result of {\bf{Step 1}}, we have
that, for every $n\in\mathbb{N}$,
$$
\begin{aligned}
F(\alpha_t^n,B_t)=&F(\alpha_0,0)+\int^t_0 \frac{\partial F}{\partial
x} (\alpha_s^n,B_s)\beta^n_s\d s +\int^t_0 \frac{\partial
F}{\partial y} (\alpha_s^n,B_s) \d B_s +\int^t_0 \frac{\partial
F}{\partial x}
(\alpha_s^n,B_s)\gamma_s\dw\d W_s\\
&-\frac12\int^t_0\frac{\partial^2 F}{\partial x^2}
(\alpha_s^n,B_s)|\gamma_s|^2\d s.
\end{aligned}
$$
Since $\alpha^n$ converges to $\alpha$ uniformly in probability on
$[0,T]$, by letting $n\to \infty$ in the above equation, we deduce
that $$\lim_{n\to\infty}\int^t_0\frac{\partial F}{\partial y}
(\alpha_s^n,B_s) \d B_s=\lim_{n\to\infty}\lim_{\ep\to
0}\int^t_0\frac{1}{\ep}\frac{\partial F}{\partial y}
(\alpha_s^n,B_s)(B_s-B_{s-\ep})\d s=\lim_{\ep\to
0}\int^t_0\frac{1}{\ep}\frac{\partial F}{\partial y}
(\alpha_s,B_s)(B_s-B_{s-\ep})\d s$$ exists, it is the Russo-Vallois
integral $\int^t_0\frac{\partial F}{\partial y} (\alpha_s,B_s)\d
B_s$ and it equals to
$$
\begin{aligned}
F(\alpha_t,B_t)-F(\alpha_0,0)-\int^t_0 \frac{\partial F}{\partial x}
(\alpha_s,B_s)\beta_s\d s -\int^t_0 \frac{\partial F}{\partial
x}(\alpha_s,B_s)\gamma_s\dw\d W_s +\frac12\int^t_0\frac{\partial^2
F}{\partial x^2} (\alpha_s,B_s)|\gamma_s|^2\d s.
\end{aligned}
$$
{\bf{Step 3}.} Finally we consider the case $F\in C^2(\R\times \R)$.
We let $\{\varphi_N\}_{N\in\N}$ be a sequence of infinitely
differentiable functions with compact support such that
$\varphi_N(x)=x$ for $\{(x_1,x_2):{\max}(|x_1|,|x_2|)\le N\},
N\in\N$. We set $F_N(x)=F(\varphi_N(x))$, so that $F_N(x)\in
C_b^2(\R\times \R)$ for every $N>0$. We notice that
$F_N(\alpha_\cdot,B_\cdot)$ and $F(\alpha_\cdot,B_\cdot)$ coincide
on the set $\Omega_N=\{ \omega\in\Omega: \sup_{0\le s\le
t}|\alpha_s|\le N,\sup_{0\le s\le t}|B_s|\le N\}$ and that
$\Omega=\bigcup_{N\ge1} \Omega_N$. Due to {\bf{Step 1}} and
{\bf{Step 2}}, for every $N$, we have
\[
\begin{aligned}
F_N(\alpha_t,B_t)=&F_N(\alpha_0,0)+\int^t_0 \frac{\partial
F_N}{\partial x} (\alpha_s,B_s)\beta_s\d s +\int^t_0 \frac{\partial
F_N}{\partial y} (\alpha_s,B_s) \d B_s
+\int^t_0 \frac{\partial F_N}{\partial x} (\alpha_s,B_s)\gamma_s\dw\d W_s\\
&-\frac12\int^t_0\frac{\partial^2 F_N}{\partial x^2}
(\alpha_s,B_s)|\gamma_s|^2\d s, \ \ t\in[0,T].
\end{aligned}
\]
Therefore, for every $N$, it holds on $\Omega_N$ that
\[
\begin{aligned}
F(\alpha_t,B_t)=&F(\alpha_0,0)+\int^t_0 \frac{\partial F}{\partial
x} (\alpha_s,B_s)\beta_s\d s +\int^t_0 \frac{\partial F}{\partial y}
(\alpha_s,B_s) \d B_s +\int^t_0 \frac{\partial F}{\partial x}
(\alpha_s,B_s)\gamma_s\dw\d W_s\\
&-\frac12\int^t_0\frac{\partial^2 F}{\partial x^2}
(\alpha_s,B_s)|\gamma_s|^2\d s,\ \ t\in[0,T].
\end{aligned}
\]
Finally, by letting $N$ tend to $+\infty$, we get the wished result.
The proof is complete. \hfill$\cajita$
\section{Doss-Sussman Transformation of Fractional Backward Doubly Stochastic Differential Equations}
\setcounter{equation}{0} In what follows, we use the following
hypotheses:
\bigskip
(H1) The functions $\sigma: \R^n\to \R^{n\times d}$ and $b:
\R^{n}\to\R^n$ are Lipschitz continuous.
\bigskip
(H2) The function $f: [0,T]\times\R^n\times\R\times\R^d\to\R$ is
Lipschitz in $(x,y,z)\in\R^n\times\R\times\R^d\to\R$ with
$|f(t,0,0,0)|\le C$ uniformly in $t\in[0,T]$, the function $g:
\R^n\to \R $ belongs to $C^3_b(\R)$ and the function $\Phi$ is
bounded.
\bigskip
We fix an arbitrary $t\in[0,T]\subset\R^+$. Let $(X_s^{t,x})_{0\leq
s \leq t}$ be the unique solution of the following stochastic
differential equation:
\begin{equation}
\label{eq:x-sde} \left\{
\begin{array}{l}
\d X_s^{t,x}=- b(X_s^{t,x})\d
s -\sigma(X_s^{t,x})\dw\d W_s , \quad s\in\left[0,t\right],\\
X_t^{t,x}=x.
\end{array}\right.
\end{equation}
Here the stochastic integral $\int^t_0\cdot\dw\d W_s$ is again
understood as the backward It\^o one. The condition (H1) guarantees
the existence and uniqueness of the solution $(X_s^{t,x})_{0\leq s
\leq t}$ in $\mathcal{M}^2(\mathbb{F}^W,[0,T];\R^n)$. Our aim is to
study the following backward doubly stochastic differential
equation:
\begin{equation}
\label{eq:x-bdsde}
U_s^{t,x}=\Phi(X_0^{t,x})+\int^s_0f(r,X_r^{t,x},
U_r^{t,x},V_r^{t,x})\d r + \int^s_0 g(U_r^{t,x})\d
B_r-\int^s_0V_r^{t,x}\dw\d W_r, \ s\in[0,t].
\end{equation}
We emphasiz that the integral with respect to the fBm $B$ is
interpreted in the Russo-Vallois sense, while the integral with
respect to the Brownian motion $W$ is the It\^o backward one. If $B$
is a standard Brownian motion, equation (\ref{eq:x-bdsde}) coincides
with the BDSDE which was first studied by Pardoux and Peng
\cite{PP2} in 1994 (apart of a time inversion).
Before we investigate the BDSDE (\ref{eq:x-bdsde}), we first give
the definition of its solution.
\begin{definition}
A solution of equation $(\ref{eq:x-bdsde})$ is a couple of processes
$(U_s^{t,x},V_s^{t,x})_{s\in[0,t]}$ such that:\\
1). $(U_s^{t,x},V_s^{t,x})_{s\in[0,t]}\in
\HH^{\infty}_t(\R)\times\HH_t^2(\R^d)$;\\
2). The Russo-Vallois integral $\int^\cdot_0g(U_r^{t,x})\d B_r$ is
well defined on $[0,t]$;\\
3). Equation $(\ref{eq:x-bdsde})$ holds $\PP$-a.s.
\end{definition}
Unlike the classical case, the lack of the semimartingale property
of the fBm $B$ gives an extra difficulty in solving BDSDE
$(\ref{eq:x-bdsde})$ directly. However, the work of Buckdahn and Ma
\cite{BM} indicates another possibility to investigate this
equation: by using the Doss-Sussman transformation. Let us develop
the idea. We denote by $\eta$ the stochastic flow which is the
unique solution of the following stochastic differential equation:
\begin{equation}
\label{eq:eta} \eta(t,y)=y+\int^t_0g(\eta(s,y))\d B_s, \ t\in[0,T],
\end{equation}
where the integral is interpreted in the sense of Russo-Vallois. The
solution of such a stochastic differential equation can be written
as $\eta(t,y)=\alpha(y,B_t)$ via Doss transformation (see, for
example, Z\"ahle \cite{Za2}), where $\alpha(y,z)$ is the solution of
the ordinary differential equation
\begin{equation}\label{eq:alpha}
\left\{
\begin{array}{l}
\frac{\partial\alpha}{\partial z} (y,z)=g(\alpha(y,z)), \ z\in\R,\\
\alpha(y,0)=y.
\end{array}\right.
\end{equation}
By the classical PDE theory we know that, for every $z\in\R$, the
mapping $y\mapsto\alpha(y,z)$ is a diffeomorphism over $\R$ and
$(y,z)\mapsto \alpha(y,z)$ is $C^2$. In particular, we can define
the $y$-inverse of $\alpha(y,z)$ and we denote it by $h(y,z)$, such
that we have $\alpha(h(y,z))=y$, $(y,z)\in\R\times\R$. Hence, it
follows that
$$
\frac{\partial\alpha}{\partial y}(h(y,z),z) \frac{\partial
h}{\partial y}(y,z)=1\,\ \mbox{and}\ \
\frac{\partial\alpha}{\partial z}(h(y,z),z)+
\frac{\partial\alpha}{\partial y}(h(y,z),z) \frac{\partial
h}{\partial z}(y,z)=0.
$$
Therefore,
\[\begin{aligned}
\frac{\partial h}{\partial
z}(y,z)&=-\left(\frac{\partial\alpha}{\partial
y}(h(y,z),z)\right)^{-1}\frac{\partial\alpha}{\partial z}(h(y,z),z)
=-\frac{\partial h}{\partial y}(y,z)\frac{\partial\alpha}{\partial
z}(h(y,z),z)=-\frac{\partial h}{\partial y}(y,z)g(y).
\end{aligned}
\]
As a direct consequence we have that also
$\eta(t,\cdot)=\alpha(\cdot,B_t):\R\to\R$ is a diffeomorphism and,
thus, we can define
$\mathcal{E}(t,y):=\eta(t,\cdot)^{-1}(y)=h(y,B_t),
(t,y)\in[0,T]\times\R$. Moreover, by the It\^o formula (Theorem
\ref{thm:ito}), we have
\[
\d\mathcal{E}(t,y)=\d h(y,B_t)=\frac{\partial }{\partial z} h(y,B_t)
\d B_t=-\frac{\partial}{\partial y}\mathcal{E}(t,y)g(y)\d B_t, \
t\in[0,T],
\]
i.e., the process $\mathcal{E}$ satisfies the following equation:
\begin{equation}
\label{eq:E} \mathcal{E}(t,y)=y-\int^t_0\frac{\partial}{\partial
y}\mathcal{E}(s,y)g(y) \d B_s, \ t\in[0,T].
\end{equation}
Furthermore, we have the following estimates for $\eta$ and
$\mathcal{E}$.
\begin{lemma}\label{prop_3.1}
There exists a constant $C>0$ depending only on the bound of $g$ and
its partial derivatives such that for $\xi=\eta, \mathcal{E}$, it
holds that, $P$-a.s., for all $(t,y)$,
\[
\begin{array}{cc}
|\xi(t,y)|\le |y|+C|B_t|,
&\exp\{-C|B_t|\}\le\lt|\frac{\partial}{\partial y}\xi\rt|
\le \exp\{C|B_t|\},\\ \lt|\frac{\partial^2}{\partial y^2}\xi\rt|\le
\exp\{C|B_t|\}, &\left|\frac{\partial^3}{\partial y^3}\xi\right|\le
\exp\{C|B_t|\}.
\end{array}
\]
\end{lemma}
\noindent\textit{Proof:} The first three estimates are similar to
those in Buckdahn and Ma \cite{BM}. So we only prove the last one.
For this end we define $\gamma(\theta,y,z)=\alpha(y,\theta z)$, for
$(\theta,y,z)\in[0,1]\times\R\times\R^n$. It follows from
(\ref{eq:alpha}) that
$$
\gamma(\theta,y,z)=y+\int^{\theta}_0\langle
g(\gamma(r,y,z)),z\rangle\d r.
$$
By differentiating this latter equation, we obtain,
\begin{equation}\label{eq:alpha_yyy}
\left\{
\begin{array}{ll}
\frac{\partial^4\gamma}{\partial \theta\partial y^3}
(\theta,y,z)=&\frac{\partial^3 g }{\partial
y^3}(\gamma(\theta,y,z))z\left( \frac{\partial \gamma}{\partial
y}(\theta,y,z)\right)^3+3 \frac{\partial^2 g }{\partial
y^2}(\gamma(\theta,y,z))z\frac{\partial \gamma}{\partial
y}(\theta,y,z)\frac{\partial^2 \gamma}{\partial
y^2}(\theta,y,z)\\&+\frac{\partial g }{\partial
y}(\gamma(\theta,y,z))z
\frac{\partial^3 \gamma}{\partial y^3}(\theta,y,z);\\
\frac{\partial^3\gamma}{\partial y^3}(0,y,z)=0,
\end{array}\right.
\end{equation}
and from the variation of parameter formula it follows that
$$
\begin{aligned}
\frac{\partial^3}{\partial y^3}\gamma(1,y,z)=\int^1_0
\exp\left\{\int^1_u \frac{\partial g }{\partial y}(\gamma(v,y,z))z\d
v\right\}&\bigg(\frac{\partial^3 g }{\partial
y^3}(\gamma(u,y,z))z\left( \frac{\partial \gamma}{\partial
y}(u,y,z)\right)^3\\
& \quad+3 \frac{\partial^2 g }{\partial
y^2}(\gamma(u,y,z))z\frac{\partial \gamma}{\partial
y}(u,y,z)\frac{\partial^2 \gamma}{\partial y^2}(u,y,z)\bigg)\d u.
\end{aligned}
$$
Thus, by using the first three estimates of this Lemma, we get
\[
\left|\frac{\partial^3}{\partial
y^3}\eta(t,y)\right|=\left|\frac{\partial^3}{\partial
y^3}\alpha(y,B_t)\right|\le \exp\{C|B_t|\}.
\]
Hence we have completed the proof. \hfill$\cajita$
Lemma \ref{prop_3.1} plays an important role in the rest of the
paper thanks to the following lemma.
\begin{lemma}\label{lemma2}
For any $C\in\R$, we have
\[
\E\lt[\exp\lt\{C\sup_{s\in[0,T]}|B_s|\rt\}\rt]<\infty.
\]
\end{lemma}
\noindent\textit{Proof:} The proof is similar as Lemma 2.4 in
\cite{Jl} (and even easier), so we omit it. \hfill$\cajita$
We denote by $\wt{\Omega}^\pr$ the subspace of $\Omega^\pr$ such
that
$\wt{\Omega}^\pr=\lt\{\omega^\pr\in\Omega^\pr:\sup_{s\in[0,T]}|B_s|<\infty\rt\}$.
It is clear that $\PP^\pr(\wt{\Omega}^\pr)=1$.
We let $(Y^{t,x}, Z^{t,x})$ be the unique solution of the following
BSDE:
\begin{equation}
\label{eq:x-BSDE}
Y_s^{t,x}=\Phi(X_0^{t,x})+\int^s_0\tilde{f}(r,X_r^{t,x},Y_r^{t,x},Z_r^{t,x})\d
r -\int^s_0 Z_r^{t,x}\dw \d W_r,
\end{equation}
where
\[\begin{aligned}
\tilde{f}(t,x,y,z)=&\frac{1}{\frac{\partial}{\partial y}\eta
(t,y)}\left\{f\lt(t,x,\eta(t,y), \frac{\partial}{\partial
y}\eta(t,y)z\rt)
+\frac12\mathrm{tr}\lt[z^T\frac{\partial^2}{\partial
y^2}\eta(t,y)z\rt]\right\}.
\end{aligned}
\]
This BSDE, studied over $\Omega=\Omega^\pr\times\Omega^{\pr\pr}$ and
driven by the Brownian motion
$W_r(\omega)=W_r(\omega^{\pr\pr})=\omega^{\pr\pr}(r)$, $r\in[0,T]$,
can be interpreted as an $\omega^\pr$-pathwise BSDE, i.e., as a BSDE
over $\Omega^{\pr\pr}$, considered for every fixed
$\omega^\pr\in\Omega^\pr.$ However, subtleties of measurability make
us preferring to consider the BSDE over $\Omega$, with respect to
the filtration $\mathbb{H}$. We point out that the coefficient
$\tilde{f}$ has a quadratic growth in $z$, while the terminal value
is bounded. BSDEs of this type have been studied by Kobylanski
\cite{Ko}. We state an existence and uniqueness result for this kind
of BSDE, but with a slight adaptation to our framework. For this we
consider a driving coefficient $G$ satisfying the following
assumptions:
\bigskip
(H3) The coefficient $G: \Omega\times[0,T]\times \R\times\R^d$ is
measurable, for every fixed $(y,z)$, progressively measurable, with
respect to the backward filtration $\mathbb{H}$ and $G$ is
continuous in $(t,y,z)$;
\bigskip
(H4) There exists some real-valued $\F_T^B$-measurable random
variable $K:\Omega^\pr\to \R$ such that $|G(t,y,z)|\le K(1+|z|^2)$.
\bigskip
(H5) There exist real-valued $\F_T^B$-measurable random variables
$C>0$, $\ep>0$, and $\F_T^B\otimes\mathcal{B}([0,T])$-measurable
functions $k,$ $l_\ep: \Omega^\pr\times[0,T]\to \R$ such that
$$
\left|\frac{\partial G}{\partial z}(t,y,z)\right|\le k(t)+C|z|, \
\mathrm{for \ all} \ (t,y,z),\ \PP- \rm{a.s.,}
$$
$$
\frac{\partial G}{\partial y} (t,y,z)\le l_\ep(t)+\ep |z|^2, \
\mathrm{for\ all}\ (t,y,z),\ \PP-\rm{a.s.}
$$
\begin{remark}\label{rmk1}
Due to the Lemmata \ref{prop_3.1} and \ref{lemma2}, the function
$\tilde{f}$ in the equation $(\ref{eq:x-BSDE})$ satisfies
$(H3)-(H5)$. In particular, $K=\exp\{C\sup_{t\in[0,T]}|B_t|\}$ for
$C\in\R^+$ appropriately chosen.
\end{remark}
Adapting the results by Kobylanski \cite{Ko} (Theorem 2.3 and
Theorem 2.6), we can state the following:
\begin{theorem}\label{thm:ko}
Let $G$ be a driver such that hypotheses $(H3)$-$(H5)$ hold and let
$\xi$ be a real-valued $\mathcal{H}_0$-measurable random variable,
which is bounded by a real-valued $\F_T^B$-measurable random
variable. Then there exists a unique solution $(Y,Z)\in
\mathcal{H}_T^\infty(\R)\times\mathcal{H}^2_T(\R^d)$ of BSDE
\begin{equation}\label{eq:BSDE}
Y_t=\xi+\int^t_0G(s,Y_s,Z_s)\d s- \int^t_0 Z_s\dw \d W_s, \
t\in[0,T].
\end{equation}
Moreover, there exists a real-valued $\F_T^B$-measurable random
variable $C$ depending only on
$\esssup_{[0,T]\times\Omega^{\pr\pr}}$
$|Y_t(\omega^\pr,\omega^{\pr\pr})|$ and $K$, such that
\[
\mathbb{E}\lt[\int^T_0|Z_s|^2\d s\big|\F_T^B\rt]\le C,\
\PP-\rm{a.s.}
\]
\end{theorem}
\begin{remark}
The conditional expectation $\E\lt[\cdot|\F_T^B\rt]$ is here
understood in the generalized sense: if $\xi$ is a nonnegative
$\HH_0$-measurable random variable,
\[
\E[\xi|\F^B_T]:=\lim_{n\to\infty}\uparrow\E[\xi\wedge n|\F_T^B](\le
\infty)
\]
is a well defined $\F_T^B$-measurable random variable. If $\xi$ is
not nonnegative we decompose $\xi=\xi^+-\xi^-, \xi^+=\max\{\xi,0\},
\xi^-=-\min\{\xi,0\}$ and we put
$\E[\xi|\F_T^B]=\E[\xi^+|\F_T^B]-\E[\xi^-|\F_T^B]$ on
$\{\min\{\E[\xi^+|\F_T^B],\E[\xi^-|\F_T^B]\}<\infty\}$.
\end{remark}
\noindent\textit{Proof of Theorem \ref{thm:ko}:} We observe that the
Brownian motion $W$ possesses the (backward) martingale
representation with respect to the backward filtration $\mathbb{H}$,
i.e., given an $\HH_0$-measurable random variable $\xi$ such that
$\E\lt[\xi^2|\F^B_T\rt]<\infty$, $\PP$-a.s., there exists a unique
process $\gamma\in \HH_T^2(\R^d)$ such that
\[
\xi=\E[\xi|\F_T^B]-\int^T_0\gamma_r\dw \d W_r, \ \PP-\rm{a.s.}
\]
This martingale representation property allows to show the existence
and uniqueness of a solution of (\ref{eq:BSDE}) when $G$ is of
linear growth and Lipschitz in $(y,z)$. Combining this with the
approach by Kobylanski \cite{Ko} allows to obtain the result stated
in Theorem \ref{thm:ko}.\hfill$\cajita$
\bigskip
By using Theorem $\ref{thm:ko}$, we are now able to characterize
more precisely the solution of BSDE (\ref{eq:x-BSDE}).
\begin{theorem}\label{thm:sol}
Under our standard assumptions on the coefficients $\sigma, b , f$
and $\Phi$, BSDE $(\ref{eq:x-BSDE})$ admits a unique solution
$(Y^{t,x}, Z^{t,x})$ in $\HH_t^\infty(\R)\times\HH_t^2(\R^d)$.
Moreover, there exists a positive increasing process $\theta\in
L^0(\mathbb{H},\R)$ such that $|Y^{t,x}_s|\le \theta_s$,
$\E\lt[\int^\tau_0|Z^{t,x}_s|^2\d s |\HH_\tau\rt]\le
\exp\{\exp\{C\sup_{s\in[0,t]}|B_s|\}\}$, $\PP$-a.s., for all
$\mathbb{H}$-stopping times $\tau$ $(0\le \tau\le t)$, where $C$ is
a constant chosen in an adequate way. Furthermore, the process
$(Y^{t,x},Z^{t,x})$ is $\mathbb{G}$-adapted.
\end{theorem}
\noindent\textit{Proof:} Due to Theorem \ref{thm:ko}, equation
(\ref{eq:x-BSDE}) has a unique solution $(Y^{t,x}, Z^{t,x})$ in
$\HH_t^\infty(\R)\times\HH_t^2(\R^d)$.
{\bf Step 1}. In order to give the estimates of $(Y^{t,x},
Z^{t,x})$, we proceed as in Lemma 5.3 in \cite{BM}. In particular,
we can show that there exists an increasing positive process
$\theta\in L^0(\mathbb{F}^B, [0,T])$ such that $\PP$-a.s.,
$|Y^{t,x}_s|\le \theta_s, 0\le s \le t \le T$. This process
$(\theta_s)$ can be chosen as the solution of the following ordinary
differential equation
\[
\frac{\d \theta_s}{\d s}=\exp\{C\sup_{0\le r\le
t}|B_r|\}(1+\theta_s);\ \theta(0)=|\Phi(X_0^{t,x})|,
\]
for some suitably chosen real constant $C$, i.e.,
$$
\theta_s=(|\Phi(X_0^{t,x})|+1)\exp\lt\{\exp\lt\{C\sup_{0\le r\le
t}|B_r|\rt\}s\rt\}-1.
$$
Indeed, for $M>0$, let $\varphi_M(y)$ be a $C^\infty$ function such
that $0\le \varphi_M\le 1$, $\varphi_M(y)=1$ for $|y|\le M$ and
$\varphi_M(y)=0$ for $|y|\ge M+1$. Defining a new function
$\tilde{f}^M$ by $\tilde{f}^M(t,x,y,z)\triangleq
\tilde{f}(t,x,y,z)\varphi_M(y)$ we see that the function
$\tilde{f}^M$ also satisfies conditions (H3)-(H5). According to
Theorem \ref{thm:ko}, there exists a unique solution
$(Y^{M,t,x},Z^{M,t,x})$ of equation (\ref{eq:x-BSDE}) with
$\tilde{f}$ being replaced with $\tilde{f}^M$. The stability result
in Kobylanski \cite{Ko} shows that, when $M\to +\infty$, there
exists a subsequence of $Y^{M,t,x}$ converging to $Y^{t,x}$
uniformly in probability. Therefore, following Buckdahn and Ma's
approach and slightly adapted, we only need to prove that
$Y^{M,t,x}_s$ is uniformly bounded by $\theta_s$. We apply the
Tanaka formula to $|Y^{M,t,x}|$ to get that
\[
\lt|Y^{M,t,x}_s\rt|=|\Phi(X_0^{t,x})|+\int^s_0\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)
\tilde{f}^M(r,X_r^{t,x},Y^{M,t,x}_r,Z^{M,t,x}_r)\d r-
\int^s_0\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)Z^{M,t,x}_r\dw \d W_r
+L_s-L_0,
\]
for a local-time-like process $L$ such that $L_t=0$ and
$L_s=\int^t_s1_{\{Y^{M,t,x}_r=0\}}\d L_r$. Now we define a new
function $\psi(y)$ by letting $\psi(y)=e^{2Ky}-1-2Ky-2K^2y^2$, for
$y>0$, and $\psi(y)=0$ for $y\le 0$, where $K$ is the bound in
Remark \ref{rmk1}. Then we apply the It\^o formula to
$\psi(\lt|Y^{M,t,x}_s\rt|-\theta_s)$ and get
\begin{equation}\label{eq:YM}
\begin{aligned}
&\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)\\=&
\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)
\lt(\tilde{f}^M\lt(r,X_r^{t,x},Y^{M,t,x}_r,Z^{M,t,x}_r\rt)
-\exp\{C\sup_{0\le u\le t}|B_u|\}(1+\theta_r)\rt)\d
r\\&-\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)Z^{M,t,x}_r\dw \d W_r
+\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\d L_r\\&
-\frac12\int^s_0\psi^{\pr\pr}\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\lt|Z^{M,t,x}_r\rt|^2\d
r
\end{aligned}
\end{equation}
The property of $\psi$ shows that
$\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\d
L_r=\int^s_0\psi^\pr\lt(-\theta_r\rt)\d L_r=0$. We also have
\[
\begin{aligned}
&\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)
\lt(\tilde{f}^M\lt(r,X_r^{t,x},Y^{M,t,x}_r,Z^{M,t,x}_r\rt)
-\exp\{C\sup_{0\le u\le t}|B_u|\}(1+\theta_r)\rt)\d r\\
\le &\int^s_0 \psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\lt(K\lt(\varphi(Y^{M,t,x}_r)\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)+K|Z^{M,t,x}_r|^2\rt)\d
r.
\end{aligned}
\]
Since $\psi^{\pr\pr}-2K\psi^\pr\ge0$, we get from equation
(\ref{eq:YM}) that
\[
\begin{aligned}
\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)\le& K
\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\lt(\varphi(Y^{M,t,x}_r)\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\d
r\\&-\int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\mathrm{sgn}\lt(Y^{M,t,x}_r\rt)Z^{M,t,x}_r\dw \d W_r.
\end{aligned}
\]
Thus, we deduce that
\[
\begin{aligned}
\E\lt[\lt.\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)\rt|\HH_s\rt]\le
&\E\lt[\lt.\ K \int^s_0\psi^\pr\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\lt(\varphi(Y^{M,t,x}_r)\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\d
r\rt|\HH_s\rt].
\end{aligned}
\]
From the definition of $\psi$ we get that
$\psi^\pr(y)=2K(\psi(y)+2K^2y^2)$. Hence, we have
\[
\begin{aligned}
&\E\lt[\lt.\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)\rt|\HH_s\rt]\\
\le&\E\lt[\lt.\ \int^s_02K^2\psi\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)
\lt(\varphi(Y^{M,t,x}_r)\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)+4K^4
\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)^3\d r\rt|\HH_s\rt].
\end{aligned}
\]
There also exists a $\tilde{K}$ such that $y^3\le\tilde{K}\psi(y)$,
for all $y\in\R$. Consequently, we get that
\[
\begin{aligned}
\E\lt[\lt.\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)\rt|\HH_s\rt]
\le2K^2(M+\|\theta\|_{\infty,[0,t]}+2K^2\tilde{K})\int^s_0
\E\lt[\lt.\psi\lt(\lt|Y^{M,t,x}_r\rt|-\theta_r\rt)\rt|\HH_s\rt]\d r.
\end{aligned}
\]
Finally, the Gronwall inequality shows that
$\psi\lt(\lt|Y^{M,t,x}_s\rt|-\theta_s\rt)=0$, for any $s\in[0,t]$,
$\PP$-a.s. Therefore, $\lt|Y^{M,t,x}_s\rt|\le\theta_s$, for any
$s\in[0,t]$, $\PP$-a.s.
{\bf Step 2}. We apply the It\^o formula to $e^{aY^{t,x}_s}$, with
$a$ being a real-valued $\F_T^B$-measurable random variable to be
determined later, and we obtain
\[
\begin{aligned}
e^{aY^{t,x}_s}&=e^{a\Phi(X_0^{t,x})}+\int^s_0ae^{aY_s}\tilde{f}(r,X^{t,x}_r,Y^{t,x}_r,
Z^{t,x}_r)\d r-\int^s_0\frac{1}{2}a^2e^{aY^{t,x}_r}|Z^{t,x}_r|^2\d
r-\int^s_0ae^{aY^{t,x}_r}Z^{t,x}_r\dw
\d W_r\\
&\le e^{a\Phi(X_0^{t,x})}+\int^s_0\left(-\frac{1}{2}a^2+|a|K\right)
e^{aY^{t,x}_r}|Y^{t,x}_r|^2\d r+\int^s_0 |a|K e^{aY^{t,x}_r}\d
r-\int^s_0ae^{aY^{t,x}_r}Z^{t,x}_r\dw \d W_r.
\end{aligned}
\]
Let
$\zeta(\omega^\pr):=\esssup_{[0,t]\times\Omega^{\pr\pr}}|Y^{t,x}_s
(\omega^\pr,\omega^{\pr\pr})|(<+\infty,
\omega^\pr\in\wt{\Omega}^\pr)$. Hence, by taking the conditional
expectations $\E[\cdot|\HH_t]$ on both sides, we deduce that for any
$\mathbb{H}$-stopping times $\tau\in[0,T]$,
\[
\begin{aligned}
&\left(\frac12a^2-|a|K\right)e^{-|a|\zeta}
\mathbb{E}\left[\int^\tau_0 |Z^{t,x}_r|^2\d
r\bigg|\HH_\tau\right]\le \left(\frac12a^2-|a|K\right)
\mathbb{E}\left[\left.\int^\tau_0 e^{aY^{t,x}_r}|Z_r|^2\d r
\rt|\HH_\tau\right]
\\
\le&\mathbb{E}\left[\left.e^{a\Phi(X_0^{t,x})}-e^{aY^{t,x}_\tau}+\int^\tau_0
|a|K e^{aY^{t,x}_r}\d s\right|\HH_\tau\right].
\end{aligned}
\]
We can choose $a=4K$ such that $\frac12a^2-|a|K=4K^2$ and we get,
keeping in mind that here $K$ is a random variable bounded by
$\exp\{C\sup_{s\in[0,t]}|B_s|\}$ ($C\in\R^+$ is a real constant, see
Remark \ref{rmk1}),
\begin{equation}\label{est-Z}
\mathbb{E}\left[\left.\int^\tau_0 |Z^{t,x}_r|^2\d r\right|
\HH_\tau\right]\le \frac{(2+4K^2t)}{4K^2}e^{8K}\le
\exp\lt\{\exp\lt\{C\sup_{s\in[0,t]}|B_s|\rt\}\rt\}.
\end{equation}
{\bf Step 3}. Let us show that the process $(Y^{t,x}, Z^{t,x})$ is
not only $\mathbb{H}$- but also $\mathbb{G}$-adapted. For this we
consider for an arbitrarily given $\tau\in[0,t]$ equation
(\ref{eq:x-BSDE}) over the time interval $[0,\tau]$:
\begin{equation}\label{x-bsde2}
Y_s^{t,x}= \Phi(X_0^{t,x}) +\int^s_0 \tilde{f}\left(r, X_r^{t,x},
Y_r^{t,x}, Z_r^{t,x}\right)\d r -\int_0^sZ_r^{t,x}\dw \d W_r, \quad
s\in\left[0,\tau\right].
\end{equation}
Let
$\mathcal{H}_t^{\tau}:=\mathcal{F}_{t,T}^W\vee\mathcal{F}_{\tau}^{B},$
$t\in[0,\tau]$. Then
$\mathbb{H}^{\tau}=\{\mathcal{H}_t^{\tau}\}_{t\in[0,\tau]}$ is a
backward Brownian filtration enlarged by a $\sigma$-algebra
generated by the fBm $B$, which is independent of the Brownian
filtration. Thus, with respect to $\mathbb{H}^\tau$, the Brownian
motion $W$ has the martingale representation property. Since
$\tilde{f}(r,x,y,z)$ is $\mathcal{G}_t$- and, hence, also
$\mathcal{H}_t^{\tau}$-measurable, $\d r$ a.e. on $[0,\tau]$, it
follows from the classical BSDE theory (or Theorem \ref{thm:ko})
that BSDE (\ref{x-bsde2}) admits a unique solution
$\left(Y^{t,x,\tau},Z^{t,x,\tau}\right) \in
\HH_\tau^\infty(\R)\times\HH_\tau^2(\R^d)$. On the other hand, also
$\left(Y^{t,x,\tau}_r,Z^{t,x,\tau}_r\right)_{r\in[0,\tau]}$ is a
solution of (\ref{eq:x-BSDE}). Hence,
$\left(Y^{t,x}_r,Z^{t,x}_r\right)=
\left(Y^{t,x,\tau}_r,Z^{t,x,\tau}_r\right)$, $\d r$ a.e., for
$t<\tau$. Consequently, $\left(Y^{t,x}_r,Z^{t,x}_r\right)$ is
$\mathcal{H}_r^\tau$- measurable, $\d r$ a.e., for $r<\tau.$
Therefore, letting $\tau\dw t$ we can deduce from the right
continuity of the filtration $\mathbb{F}^B$ that
$\left(Y^{t,x},Z^{t,x}\right)$ is
$\mathbb{G}$-adapted.\hfill$\cajita$
\begin{remark}
We remind the reader that the bound we get in $(\ref{est-Z})$ is
only $\PP$-a.s. finite, but not square-integrable. As a matter of
fact, it is hard to prove directly that $Z^{t,x}$ is a
square-integrable process, which constitutes the main reason that we
use instead the space $\HH_t^2(\R^d)$. Hence, the major difference
between our work and Buckdahn and Ma \cite{BM} is: In the classical
case, a priori we can solve the BDSDE in the first step to get the
square integrability of $Z$, but in the fractional case, there is
not a direct way to solve the BDSDE.
\end{remark}
Now we are ready to give the main result of this section by linking
the BSDEs (\ref{eq:x-bdsde}) and (\ref{eq:x-BSDE}) with the help of
the Doss-Sussman transformation.
\begin{theorem}\label{thm:ds}
Let us define a new pair of processes $(U^{t,x}, V^{t,x})$ by
$$
U_s^{t,x}=\eta(s, Y_s^{t,x}), \ \ V_s^{t,x}=\frac{\partial}{\partial
y}\eta(s, Y_s^{t,x})Z_s^{t,x},
$$
where $(Y^{t,x},Z^{t,x})$ is the solution of BSDE
$(\ref{eq:x-BSDE})$. Then $(U^{t,x}, V^{t,x})\in
\HH_t^{\infty}(\R)\times\HH_t^2(\R^d)$ is the solution of BDSDE
$(\ref{eq:x-bdsde})$.
\end{theorem}
\begin{remark}
The above theorem can be considered as a counterpart of Theorem 3.9
in Jing and Le\'on \cite{Jl} for the semi-linear case when $H<1/2$.
However, since we use here a different Hurst parameter $H$ and a
different type of stochastic integral with respect to fBm $B$, the
above theorem obviously does not cover the result in \cite{Jl}.
\end{remark}
\noindent\textit{Proof of Theorem \ref{thm:ds}:} The fact that
$(U^{t,x}, V^{t,x})\in \HH_t^{\infty}(\R)\times\HH_t^2(\R^d)$
follows directly from Theorem \ref{thm:sol} and Lemma
\ref{prop_3.1}. In order to prove the remaining part of the theorem,
we just have to apply the It\^o formula (Theorem \ref{thm:ito}) to
$\alpha(Y_s^{t,x},B_s)$, noticing $\alpha(Y_s^{t,x},B_s)=\eta(s,
Y_s^{t,x})$, to obtain that, for $(s,x)\in[0,t]\times\R^d$, the
Russso-Vallois integral $\int^s_0g(U^{t,x}_r)\d B_r$ exists and
\begin{equation}
\begin{aligned}
U_s^{t,x}=&\Phi_0(X_0^{t,x}) +\int^s_0 \frac{\partial}{\partial
y}\alpha( Y_r^{t,x},B_r)\bigg[\lt({\frac{\partial}{\partial y}\eta
(r, Y_r^{t,x})}\rt)^{-1} \bigg\{f\left(r,X_r^{t,x},\eta(s,
Y_r^{t,x}), \frac{\partial}{\partial y}\eta(r,
Y_r^{t,x})Z_r^{t,x}\right)
\\
& +\frac12\mathrm{tr}\lt[(Z^{t,x}_r)^T\frac{\partial^2}{\partial
y^2}\eta(r, Y_r^{t,x})Z_r^{t,x}\rt]\bigg\}\d r-Z_r^{t,x}\dw\d
W_r\bigg]
-\frac12\int^s_0\mathrm{tr}\lt[(Z^{t,x}_r)^T\frac{\partial^2}{\partial
y^2}\eta(r, Y_r^{t,x})Z_r^{t,x}\rt]\d
r\\
&+\int^s_0 \frac{\partial}{\partial z}\alpha( Y_r^{t,x},B_r) \d B_r.
\end{aligned}
\end{equation}
Consequently,
\begin{equation}
\begin{aligned}
U_s^{t,x}=\Phi(X_0^{t,x})+\int^s_0
f(r,X_r^{t,x},U_r^{t,x},V_r^{t,x})\d r + \int^s_0 g(U_r^{t,x}) \d
B_r-\int^s_0V_r^{t,x}\dw \d W_r.
\end{aligned}
\end{equation}
The proof is complete.\hfill$\cajita$
Now we close this section by a simple example to illustrate the idea
of the above procedure.
\begin{example}
Let us consider the linear case $($ for simplicity of notations we
omit the superscript $(t,x)$$)$:
\begin{equation}\label{eq:example}
\left\{
\begin{array}{l}
\d U_s= U_s\d B_s+f(U_s,V_s)\d s +V_s\dw \d W_s, \ s\in[0,t];\\
U_0=\Phi(X_0^{t,x}).
\end{array} \right.
\end{equation}
It is elementary to show that $\alpha(y,z)=ye^z$. Hence, the
solution $(U,V)$ of equation $(\ref{eq:example})$ is given by
$U_s=Y_se^{B_s}$ and $V_s=Z_se^{B_s}$, where $(Y,Z)$ is the solution
of
\begin{equation}\label{eq:example1}
\left\{
\begin{array}{l}
\d Y_s= \hat{f}(Y_s,Z_s)\d s +Z_s\dw \d W_s, \ s\in[0,t];\\
Y_0=\Phi(X_0^{t,x}).
\end{array} \right.
\end{equation}
and the function $\hat{f}$ is defined by
$\hat{f}(y,x)=e^{-B_t}f(ye^{B_t},ze^{B_t})$. Obviously $\hat{f}$ is
Lipschitz in $(y,z)$ as far as $f$ is Lipschitz. By following a
classical Malliavin calculus method $($see, e.g., Pardoux and Peng
\cite{PP2}$)$, we can get that $Z$ is $\PP$-a.s. uniformly bounded
and, thus, the process $\{Y_s,s\in[0,t]\}$ is $(1/2-\ep)$-H\"older
continuous in $s$, for all $\ep\in (0,1/2)$. Therefore, we have, for
every $r,s\in [0,t]$,
\[
\begin{aligned}
&|U_r-U_s|=|Y_re^{B_r}-Y_se^{B_s}|\le
\|Y\|_{\infty}|e^{B_r}-e^{B_s}|+ e^{|B_s|}|Y_r-Y_s|\\
\le&\exp\lt\{\sup_{0\le u\le
t}|B_u|\rt\}\|Y\|_{\infty}C(\omega^\pr)|r-s|^\alpha+\exp\lt\{\sup_{0\le
u\le t}|B_u|\rt\}C(\omega^\pr)|r-s|^{1/2-\ep}, \rm{for}\ \rm{a.a.}
\omega^\pr\in \Omega^\pr,
\end{aligned}
\]
where we can choose $\alpha$ to be in $(1/2,H)$. That is to say, we
can choose $0< \ep< \alpha-1/2$ and get that the process $\{U_s,
s\in [0,t]\}$ has $\alpha$-H\"older continuous paths. Hence, instead
of using the Russo-Vallois integral with respect to the fractional
Brownian motion in equation $(\ref{eq:example})$ we can use the
classical Young integral. Furthermore, thanks to Proposition
\ref{prop_rv1}, these two integrals coincide.
\end{example}
\begin{remark}\label{rmk2}
\em{In the latter example, the Young integral and the Russo-Vallois
integral coincide. Unfortunately, the H\"older continuity seems to
be very hard to deduce in the general, nonlinear case. As a
consequence, we have to work with the more general Russo-Vallois
integral.}
\end{remark}
\section{Associated Stochastic Partial Differential Equations}
\setcounter{equation}{0}
In this section we discuss briefly the relationship between an
associated SPDE and a PDE with stochastic coeffients. For
simplicity, we only show the relationship for the case of classical
solutions. For a complete discussion of the case of (stochastic)
viscosity solutions, one can proceed by adapting the approaches in
Buckdahn and Ma \cite{BM} and Jing and Le\'on \cite{Jl}.
Let $\mathscr{L}$ be the second order elliptic differential
operator:
\[
\mathscr{L}=\frac12\sum_{i,j=1}^n (\sigma\sigma^T)_{ij}(x)
\frac{\partial^2 }{\partial x_i \partial x_j}+ \sum_{i=1}^n b_i(x)
\frac{\partial }{\partial x_i},
\]
which means that it is the infinitesimal generator of the Markovian
process $\{X^{t,x}_s, s\in[0,t]\}$ defined by equation
(\ref{eq:x-sde}).
Our aim is to study the following semilinear SPDE driven by the fBm
$B$:
\begin{equation}\label{spde}
\left\{
\begin{array}{lc}
\d v(t,x) = \left[\mathscr{L} v (t,x) -
{f}(t,x,v(t,x),\sigma(x)^T\frac{\partial}{\partial x}v(t,x))\rt]\d t
+ g(v(t,x))\d B_t,& (t,x)\in (0,T)\times
\R^n;\\
v(0,x)=\Phi(x),&x\in\R^n.
\end{array}\right.
\end{equation}
In the case that $B$ is a Brownian motion (an fBm with Hurst
parameter $H=1/2$), it is well known (see, Pardoux and Peng
\cite{PP2}, Buckdahn and Ma \cite{BM}) that the random field
$v(t,x):=U^{t,x}_t$ solves $(\ref{spde})$ (in the viscosity sense if
the coefficients are Lipschitz, and in the classical sense if the
coefficients are $C^3_b$), where $U^{t,x}$ is the solution of BDSDE
(\ref{eq:x-bdsde}). Thus, it is natural to raise the following
question: Can we also solve the SPDE (\ref{spde}) driven by the fBm
$B$, by studying the properties of the solution of its associated
BDSDE (\ref{eq:x-bdsde})? The answer is positive. Indeed, by
applying the Doss-Sussman transformation, we will show that the PDE:
\begin{equation}\label{w-pde}
\left\{
\begin{array}{cllc}
\d u(t,x)&=&\lt(\mathscr{L} u (t,x) -
\tilde{f}(t,x,u(t,x),\sigma(x)^T\frac{\partial}{\partial
x}u(t,x))\rt)\d t,& (t,x)\in (0,T)\times
\R^n;\\
u(0,x)&=&\Phi(x),&x\in\R^n,
\end{array}\right.
\end{equation}
is transformed into SPDE (\ref{spde}), where we recall that
$$
\tilde{f}(t,x,y,z)={\lt(\frac{\partial}{\partial y} \eta
(t,y)\rt)^{-1}}\left\{f(t,x,\eta(t,y), \frac{\partial}{\partial
y}\eta(t,y)z) +\frac12\mathrm{Tr}\lt[z\frac{\partial^2}{\partial
y^2}\eta(t,y)z^T\rt]\right\}.
$$
We also observe that following a similar argument as in Kobylanski
\cite{Ko}, under some smoothness assumptions, $u(t,x):=Y^{t,x}_t,
(t,x)\in[0,T]\times\R^n$ is the solution of equation (\ref{w-pde}),
where $Y^{t,x}$ is the solution of BSDE (\ref{eq:x-BSDE}).
First we give the definition for the classical solutions of
equations (\ref{spde}) and (\ref{w-pde}).
\begin{definition}
We say a stochastic field $w: \Omega^\pr\times[0,T]\times\R^n\to\R$
is a classical solution of equation $(\ref{spde})$ $($resp.,
$(\ref{w-pde}))$, if $w\in C^{0,2}_{\mathbb{F}^B}$
\footnote{$C^{0,2}_{\mathbb{F}^B}$ is the space of
${\mathbb{F}^B}$-adapted processes $u(t,x)$ which are continuous in
$t$ and $C^2$ in $x$.} and satisfies equation $(\ref{spde})$
$($resp., $(\ref{w-pde}))$.
\end{definition}
We have the following proposition.
\begin{proposition}\label{prop_u}
Suppose that $u$ is a classical solution of equation
$(\ref{w-pde})$, then
$\hat{u}(t,x)\triangleq\eta(t,u(t,x))=\alpha(u(t,x),B_t)$ is a
classical solution of SPDE $(\ref{spde})$. The converse holds also
true: Every classical solution $\hat{u}$ of equation $(\ref{spde})$
defines a classical solution $u(t,x)=\mathcal{E}(t,\hat{u}(t,x))$ of
equation $(\ref{w-pde})$.
\end{proposition}
\noindent\textit{Proof:} The claim that $\hat{u}(t,x)\in
C^{0,2}_{\mathbb{F}^B}$ follows from the regularity property of the
functions $\alpha$ and $u$ and the fact that $u$ is
$\mathbb{F}^B$-adapted. Moreover, we first observe that
$\hat{u}(0,x)=\eta(0,u(0,x))=u(0,x)=\Phi(x)$, and we apply the It\^o
formula to $\hat{u}(t,x)$ to obtain:
\begin{equation}\label{eq:hatu}
\begin{aligned}
\d \hat{u}(t,x) =& \frac{\partial}{\partial
y}\alpha(u(t,x),B_t)\lt(\mathscr{L} u (t,x) -
\tilde{f}\lt(t,x,u(t,x),\sigma(x)^T\frac{\partial}{\partial
x}u(t,x)\rt)\rt)\d t
+\frac{\partial}{\partial z}\alpha(u(t,x),B_t)\d B_t\\
=& \Bigg[\frac{\partial}{\partial y}\alpha(u(t,x),B_t)\mathscr{L} u
(t,x)- f\lt(t,x,\eta(t,u(t,x)),\frac{\partial}{\partial
y}\eta(t,u(t,x))\sigma(x)^T\frac{\partial}{\partial x}u(t,x)\rt)\\
&\quad-\frac12\mathrm{Tr}\lt[\sigma(x)^T\frac{\partial}{\partial
x}u(t,x)\frac{\partial^2}{\partial y^2}
\eta(t,u(t,x))\frac{\partial}{\partial
x}u(t,x)^T\sigma(x)\rt]\Bigg]\d t+ \frac{\partial}{\partial
z}\alpha(u(t,x),B_t)\d B_t.
\end{aligned}
\end{equation}
We notice that
$$
\frac{\partial}{\partial x}\hat{u}(t,x)=\frac{\partial}{\partial
x}[\alpha(u(t,x),B_t)]=\lt(\frac{\partial}{\partial
y}\alpha\rt)(u(t,x),B_t)\frac{\partial}{\partial x}u(t,x),
$$
\[
\begin{aligned}
\frac{\partial^2}{\partial x^2}\hat{u}(t,x)
=\frac{\partial}{\partial x}u(t,x)\lt(\frac{\partial^2}{\partial
y^2}\alpha\rt)(u(t,x),B_t)\lt(\frac{\partial}{\partial
x}u(t,x)\rt)^T+ \lt(\frac{\partial}{\partial
y}\alpha\rt)(u(t,x),B_t)\frac{\partial^2}{\partial x^2}u(t,x).
\end{aligned}
\]
Thus, by substituting the above relations into (\ref{eq:hatu}), we
can simplify the equation (\ref{eq:hatu}) and we get
\begin{equation}
\d \hat{u}(t,x)=\lt[\mathscr{L} \hat{u}
(t,x)-f\lt(t,x,\hat{u}(t,x),\sigma(x)^T\frac{\partial}{\partial
x}\hat{u}(t,x)\rt)\rt]\d t+ g(\hat{u}(x,t))\d B_t.
\end{equation}
Consequently, $\hat{u}$ is a solution of SPDE (\ref{spde}). The
proof of the converse direction is analogous. The proof is complete.
\hfill$\cajita$
In order to summarize, we have constructed a solution of SPDE
(\ref{spde}) with the help of the fractional BDSDE
(\ref{eq:x-bdsde}), by passing through the quadratic BSDE
(\ref{eq:x-BSDE}) and the associated PDE (\ref{w-pde}) with random
coefficients:
\[
\begin{array}{cccc}
&fract.\ BDSDE \ (\ref{eq:x-bdsde})&\stackrel{Thm. \
\ref{thm:ds}}{\Longrightarrow} &quadratic \ BSDE \ (\ref{eq:x-BSDE})\\
&&&\Downarrow \\
&fract.\ SPDE \ (\ref{spde})&\stackrel{ Prop.\
\ref{prop_u}}\Longleftrightarrow &PDE \ (\ref{w-pde}) \ (with \
random\ coefficients).
\end{array}
\]
Proposition \ref{prop_u} shows that, in the case of a classical
solution, the Doss-Sussman transformation establishes a link between
PDE (\ref{w-pde}) and SPDE (\ref{spde}). This motivates us to give
the following definition of the stochastic viscosity solution. For
the cases $H\in(0,1/2]$, the reader is referred to Buckdahn and Ma
\cite{BM}, Jing and Le\'on \cite{Jl}. For the classical definition
of the viscosity solution, we refer to Crandall et al. \cite{CIL}.
\begin{definition}
A continuous random field $\hat{u}: \Omega^{\pr}\times[0,T] \times
\R^n
\to\R$ is called a $($stochastic$)$ viscosity solution
of equation $(\ref{spde})$ if and only if
$u(t,x)=\mathcal{E}(t,\hat{u}(t,x)), (t,x)\in[0,T]\times\R^n$ is the
viscosity solution of equation $(\ref{w-pde})$.
\end{definition}
By following a similar argument as that developed in the proof of
Theorem 4.9 in Jing and Le\'on \cite{Jl}, our preceding discussion
leads to the following theorem:
\begin{theorem}
The stochastic field $\hat{u}: \Omega^\pr\times [0,T]\times \R^n\to
\R$ defined by $\hat{u}(t,x)\triangleq \alpha (Y_t^{t,x},B_t)$ is a
$($stochastic$)$ viscosity solution of SPDE $(\ref{spde})$.
\end{theorem}
\section*{Acknowledgements}
The author thanks Professor Rainer Buckdahn for his valuable
discussions and advice.
|
\section{Introduction}
In 2007 Edwards proposed a new normal form for elliptic curves over a field $k$ of characteristic $\ne 2$~\cite{edwards}, namely:
\begin{equation}\label{original}
E_{a}(k) : x^2 + y^2 = a^2(1 + x^2 y^2),
\end{equation}
for $a^5 \ne a$. Bernstein and Lange generalised Edwards' form to incorporate curves of the form
$$E(k):x^2 + y^2 = a^2(1+dx^2y^2),$$
which is elliptic if $ad(1-da^4) \ne 0$~\cite{bl}. All curves in the
Bernstein-Lange form are isomorphic to curves of the following form, referred to as
Edwards curves:
\begin{equation}\label{edwards}
E_d(k) : x^2 + y^2 = 1 + d x^2y^2.
\end{equation}
Edwards curves over finite fields are of great interest in cryptography since the addition and
doubling formulae are: {\em unified}, which protects against some side-channel
attacks~\cite[Chapters 4 and 5]{ecc2}; {\em complete} when $d$ is a non-square,
which means the addition formulae work for all input points; and are the most
efficient in the literature.
Bernstein {\em et al. } have also considered {\em twisted} Edwards curves~\cite{twisted}:
\begin{equation}\label{twisted}
E_{a,d}(k):ax^2 + y^2 = 1 + dx^2y^2,
\end{equation}
which includes more curves over finite fields than does Edwards curves.
Rezaeian and Shparlinski have computed the exact number of distinct
curves of the form~(\ref{original}) and~(\ref{edwards}) over a finite
field ${\bbbf}_q$ of characteristic $> 2$, up to isomorphism over the algebraic
closure of ${\bbbf}_q$~\cite{shparlinski}. However they state that counting the number of
distinct isogeny classes over ${\bbbf}_q$ for these curves is a very natural and challenging question.
In this paper we answer this question fully for fields of characteristic $>2$.
Our starting point is interesting in that it was
serendipitous, beginning with an incidental empirical observation.
When searching for suitable parameters for elliptic curve cryptography, for curves of the form~(\ref{edwards}),
we observed that over a finite field ${\bbbf}_p$ with $p \equiv 1 \pmod{4}$,
it (empirically) holds that
\[
\#E_d({\bbbf}_p) = \#E_{1-d}({\bbbf}_p),
\]
and hence by Tate's theorem~\cite{tate}, $E_d$ and $E_{1-d}$ should be isogenous over ${\bbbf}_p$.
In the course of proving the above observation using character sum identities, we discovered that the Edwards curve $E_d$
is isogenous to the Legendre curve:
\begin{equation}\label{legendre}
L_d({\FF_q}): y^2 = x(x-1)(x-d).
\end{equation}
With explicit computation one sees that this isogeny has degree two,
and so $E_d$ inherits a set of $4$-isogenies from the well-known set of isomorphisms of $L_d$,
each as the composition of the $2$-isogeny to $L_d$, an isomorphism of $L_{d}$
to $L_{d'}$, and the dual of the $2$-isogeny from $E_{d'}$ to $L_{d'}$.
In particular $E_d/{\bbbf}_p$ is $4$-isogenous to $E_{1-d}/{\bbbf}_p$ for $p \equiv 1 \pmod{4}$.
More generally, for $E_d$ over any finite field ${\bbbf}_q$ one obtains $4$-isogenies
to $E_{1-d}$, $E_{1/d}$, $E_{1-1/d}$, $E_{1/(1-d)}$ and $E_{d/(d-1)}$,
being defined over ${\bbbf}_q$ or ${\bbbf}_{q^2}$ depending on the quadratic character
of $-1,d$ and $1-d$ in ${\bbbf}_q$.
We later learned that the above $2$-isogeny is merely a special case of Theorem 5.1 of~\cite{twisted}, which
states that any elliptic curve with three ${\bbbf}_q$-rational $2$-torsion
points is $2$-isogenous to a twisted Edwards curve of the form~(\ref{twisted}). However
the explicit connection with the Legendre curve and the consequent ramifications contained herein
has --- to the best of our knowledge --- not been made before.
Using the explicit connection with Legendre curves, counting the number of isogeny classes of Edwards curves
is straightforward; we use a recent result due to Katz~\cite{Katz-ratio}, who studied the isogeny
classes of Legendre curves. In doing so, we also count the number of supersingular parameters $d$ for
Edwards curves. We then prove the existence of complete Edwards curves in every isogeny class, providing
formulae for the proportion of $d \in {\bbbf}_q \setminus \{0,1\}$ for
which $L_d$ --- and hence $E_d$ --- is complete, relative to the total number of $d$ in each
isogeny class. This total be computed via a Deuring-style class number formula derived by
Katz~\cite{Katz-ratio}, and hence for a given trace one can compute the number of
complete Edwards curve parameters $d$.
We also address the distribution of original Edwards curves~(\ref{original})
amongst the isogeny classes of Edwards curves. For $q \equiv -1 \pmod{4}$ this
follows from our result on complete Edwards curves, but for $q \equiv 1 \pmod{4}$ we express the proportion of
such curves in a given isogeny class using a set of remarkable ratio results due to
Katz~\cite{Katz-ratio}. Whilst we believe our results may be proven succinctly
using a variation of Katz's approach, our arguments for the proportion of complete and original
Edwards curves rely only on explicit bijections between sets of curves of
different parameter types, and are thus entirely elementary.
\vspace{3mm}
\noindent {\em Notation:} For two elliptic curves over a field $k$, we write $E \sim E'$ when $E$ is isogenous to $E'$ over $k$, and
$E \cong E'$ when $E$ is isomorphic to $E'$ over the algebraic closure of $k$. Throughout the paper, ${\bbbf}_p$
refers to a finite field of prime cardinality $p$ and ${\bbbf}_q$ to an extension
field of cardinality $q=p^m$, where $m \geq 1$. Also,
if the field of definition of a curve or map is not specified, it is assumed
to be a field of characteristic $\ne 2$.
\section{A point counting proof of $E_d({\FF_q}) \sim L_d({\FF_q})$}\label{pointcount}
It is well known that the elliptic integral
$$
\int\frac{p(x)}{\sqrt{q(x)}}dx,
$$
where $p(x)\in{\bbbr}(x)$ is a rational function and $q(x)\in{\bbbr}[x]$ is a quartic
polynomial, can be reduced to
$$
\int\frac{p_1(x)}{\sqrt{q_1(x)}}dx
$$
for a rational function $p_1(x)\in{\bbbr}(x)$ and a cubic polynomial $q_1(x)\in{\bbbr}[x]$ provided that
one knows one of the roots of $q(x)$~\cite[Chapter 8]{integral}.
The finite field analogue of this fact is the following result of
Williams~\cite{KSW}.
\begin{lemma}\label{q-c-s}~\cite{KSW} Let $q$ be an odd prime power and let ${\bbbf}_q$ denote
the finite field with $q$ elements. Suppose that $F(x)$ is a complex valued function
from ${\bbbf}_q$ to ${\bbbc}$ and also let $\chi_2(\cdot)$ denote the quadratic character of ${\bbbf}_q$.
Also let $Z$ denote the zero set of $a_2x^2+b_2x+c_2$. Then
\begin{eqnarray}
\sum_{x\in {\bbbf}_q \setminus Z} F\left(\frac{a_1x^2+b_1x+c_1}{a_2x^2+b_2x+c_2}\right)&=&
\sum_{x\in {\bbbf}_q}
\chi_2(Dx^2+\Delta x+d)F(x)\\\nonumber&&+\sum_{x\in {\bbbf}_q} F(x)-\left\{
\begin{array}{clcl}
F\left(\frac{a_1}{a_2}\right), & \;\;\;\mbox{if $a_2\neq 0$}, \\\\
0,&\;\;\;\mbox{otherwise},\\
\end{array}\right.
\end{eqnarray}
where $a_1,b_1,c_1,a_2,b_2,c_2\in{\bbbf}_q$,
\begin{equation}
D=b_2^2-4a_2c_2,\; \Delta=4a_1c_2-2b_1b_2+4a_2c_1,\; d=b_1^2-4a_1c_1,
\end{equation}
and
$$\Delta^2-4dD\neq 0.$$
\end{lemma}
In the following we use the lemma above to show that $E_d({\FF_q})$ is isogenous to
$L_d({\FF_q})$. First notice that the given singular model for Edwards
curves~(\ref{edwards}) has two points at infinity which are singular
and no affine singular points, and resolving the singularities results in four points which
are defined over ${\FF_q}$ if and only if $d$ is a quadratic residue in
${\FF_q}$~\cite{bl}. Thus the non-singular model of $E_d({\FF_q})$ has $2+2\chi_2(d)$
points more than the singular model of $E_d({\FF_q})$, and
hence if we rewrite the curve equation of $E_d$ as
\begin{equation}\label{frac}
E_d({\FF_q}): y^2=\frac{x^2-1}{dx^2-1},
\end{equation}
then
\begin{eqnarray}\label{Edwards-order}
\#E_d({\FF_q})&=&2+2\chi_2(d)+\sum_{x\in{\bbbf}_q,\; x\ne \pm d^{1/2}}
(1+\chi_2\left(\frac{x^2-1}{dx^2-1}\right)) \nonumber\\
&=&2+2\chi_2(d)+q-(1+\chi_2(d))+
\sum_{x\in{\bbbf}_q,\; x\ne \pm d^{1/2}}\chi_2\left(\frac{x^2-1}{dx^2-1}\right)\nonumber\\
&=&q+1+\chi_2(d)+
\sum_{x\in{\bbbf}_q,\; x\ne \pm d^{1/2}}\chi_2\left(\frac{x^2-1}{dx^2-1}\right).
\end{eqnarray}
Now on the one hand by applying Lemma~\ref{q-c-s} with $F(x)=\chi_2(x)$, we get
\begin{eqnarray}\label{Ed-Leg}
\sum_{x\in{\bbbf}_q,\; x\ne \pm d^{1/2}}\chi_2\left(\frac{x^2-1}{dx^2-1}\right)&=&
\sum_{x\in {\bbbf}_q}
\chi_2(4dx^2-(4+4d)x+4)\chi_2(x) \nonumber \\
\nonumber &+& \sum_{x\in {\bbbf}_q} \chi_2(x)-\chi_2(d)\\
&=&\sum_{x\in {\bbbf}_q} \chi_2((x-1)(dx-1))\chi_2(x)-\chi_2(d)\nonumber\\
&=&\sum_{x\in {\bbbf}_q} \chi_2(x(x-1)(x-d))-\chi_2(d),
\end{eqnarray}
and on the other hand we have
\begin{equation}\label{Legendre-order}
\#L_d({\FF_q})=q+1+\sum_{x\in{\bbbf}_q}\chi_2(x(x-1)(x-d)),
\end{equation}
where $-\sum_{x\in{\bbbf}_q}\chi_2(x(1-x)(x-d))$ is the trace of the Frobenius endomorphism.
Thus comparing~(\ref{Edwards-order}),~(\ref{Ed-Leg}),~(\ref{Legendre-order}) we have:
\begin{theorem}\label{mainisog}
The Edwards curve $E_d({\FF_q})$ and Legendre curve $L_d({\FF_q})$ are isogenous.
\end{theorem}
Lemma~\ref{q-c-s} can be viewed as a means of establishing isogeny relations between curves
defined by relations such as
\begin{equation}\label{quadratio}
y^2 = \frac{a_1x^2+b_1x+c_1}{a_2x^2+b_2x+c_2},
\end{equation}
and curves defined by $y^2=x(Dx^2+\Delta x+d)$. In \S\ref{quad} we show how to derive
an addition law for curves of the form~(\ref{quadratio}) and prove results similar to those presented in the
intervening sections.
\section{$4$-isogenies of $E_d$}
In this section we detail how to compute explicit $4$-isogenies for $E_d$,
starting with the $2$-isogeny from $E_d$ to $L_d$ and its dual. We then detail
the well-known isomorphisms of $L_d$ and compose these maps to form the desired $4$-isogenies.
\subsection{Explicit $2$-isogeny $\psi_d:E_d \rightarrow L_d$}
We now derive a $2$-isogeny from $E_d$ to $L_d$, as presented in
the following result.
\begin{theorem}\label{2isogeny}
Let $(x,y) \in E_d$. Then $\psi_d:E_d \rightarrow L_d$
\[
(x,y) \mapsto \left( \frac{1}{x^2}, \frac{y(d-1)}{x(1-y^2)}\right).
\]
is a 2-isogeny.
The dual of $\psi_d$ is $\widehat{\psi}_d:L_d \rightarrow E_d:$
\[
(x,y) \mapsto \left( \frac{2{y}}{d-{x}^2}, \frac{{y}^2 - {x}^2(1-d)}{{y}^2 + {x}^2(1-d)}\right).
\]
\end{theorem}
Note that $\psi_d$ is defined on all points of $E_d$ except the kernel elements
$(0,\pm1)$, which map to $\mathcal{O} \in L_d$.
\begin{proof}
One has the following birational transformation $\tau$
\[
\tau(x,y) = \left( (1-d) \frac{1+y}{1-y}, (1-d) \frac{2(1+y)}{x(1-y)}\right),
\]
from $E_d$ to the Weierstrass curve
\[
W_d : y^2 = x^3 + 2(1+d)x^2 + (1-d)^2 x,
\]
with inverse
\[
\tau^{-1}(x,y) = \left( \frac{2x}{y}, \frac{x - (1-d)}{x + (1-d)}\right).
\]
While $\tau$ is not defined for the points $(0,\pm 1) \in E_d$, one obtains an
everywhere-defined isomorphism between the respective desingularized
projective models by sending $(0,1)$ to $\mathcal{O} \in W_d$ and $(0,-1)$ to $(0,0)$. Similarly,
$\tau^{-1}$ is not defined at points $(x,y) \in W_d$ satisying $y(x+1-d) = 0$, but if $d$ is a square the points other
than $(0,0)$ map to points of order $2$ and $4$ at infinity on the
desingularisation of $E_d$ (see the discussion on exceptional points after
Theorem 3.2 of~\cite{twisted}).
The $2$-isogeny used in the proof of Theorem 5.1 of~\cite{twisted} now maps
$W_d$ directly to $L_d$ via
\[
\phi_d(x,y) = \left( \frac{y^2}{4x^2}, \frac{y((1-d)^2 - x^2)}{8x^2}\right),
\]
with dual
\[
\widehat{\phi}_d(x,y) = \left( \frac{y^2}{x^2}, \frac{y(d - x^2)}{x^2}\right).
\]
One can verify that the compositions $\phi_d \circ \tau$ and $\widehat{\tau} \circ \widehat{\phi}_d$ give the
stated $\psi_d$ and $\widehat{\psi}_d$ respectively.
\end{proof}
\subsection{Isomorphisms of $L_d$}
The set of isomorphisms of $L_d$ are induced by the two involutions
$\sigma_1(d) = 1-d$ and $\sigma_{2}(d)=1/d$, which induce the following maps from
$L_d$ to $L_{1-d}$ and $L_{1/d}$ respectively:
\begin{eqnarray}
\label{s1} &\sigma_1:& L_d \longrightarrow L_{1-d}: (x,y) \mapsto (1-x,\sqrt{-1}y),\\
\label{s2} &\sigma_2:& L_d \longrightarrow L_{1/d}: (x,y) \mapsto (x/d,y/d^{3/2}).
\end{eqnarray}
As transformations acting on a given field, the group generated by $\sigma_1,\sigma_2$ is:
\[
H = \{1,\sigma_1,\sigma_2, \sigma_1\sigma_2,\sigma_2\sigma_1, \sigma_1\sigma_2\sigma_1\},
\]
which is isomorphic to the symmetric group $S_3$. The orbit of $d\neq 0,1$ under the action of
$H$ is
\begin{equation}\label{isovalues}
\Big\{ d, 1-d, \frac{1}{d}, 1-\frac{1}{d}, \frac{1}{1-d}, \frac{d}{d-1} \Big\}
\end{equation}
which has 6 distinct elements provided that
$d$ is
not a root of $d^2 - d + 1=0$ or $(d+1)(d-2)(2d-1)=0$.
Hence we have isomorphisms between each pair of $L_d$, $L_{1-d}$,
$L_{1/d}$, $L_{1- 1/d}$, $L_{1/(1-d)}$ and $L_{d/(d-1)}$. For
completeness we give here the remaining three isomorphisms from $L_d$ to
$L_{\sigma(d)}$ not listed in~(\ref{s1}),(\ref{s2}):
\begin{eqnarray}
\label{s12} &\sigma_1\sigma_2:& L_d \longrightarrow L_{1-\frac{1}{d}}: (x,y) \mapsto (1-x/d,\sqrt{-1}y/d^{3/2}),\\
\label{s21} &\sigma_2\sigma_1:& L_d \longrightarrow L_{\frac{1}{1-d}}: (x,y) \mapsto \left(\frac{1-x}{1-d},\frac{\sqrt{-1}y}{(1-d)^{3/2}}\right),\\
\label{s121} &\sigma_1\sigma_2\sigma_1:& L_d \longrightarrow L_{\frac{d}{d-1}}: (x,y) \mapsto \left(\frac{x-d}{1-d},-\frac{y}{(1-d)^{3/2}}\right).
\end{eqnarray}
\subsection{$4$-isogenies of $E_d$ to $E_{\sigma(d)}$}
Let $\sigma \in H$. Then $\omega_{\sigma(d)}: E_d \rightarrow E_{\sigma(d)}$ is obtained as the following composition:
\[
\omega_{\sigma(d)} = \widehat{\psi}_{\sigma(d)} \circ \sigma \circ \psi_d.
\]
The $2$-isogeny $\widehat{\psi}_{\sigma(d)}$ can be obtained by taking $\widehat{\psi}_d$ and substituting $\sigma(d)$ for $d$.
We do not write down all possible $4$-isogenies but note that whether each is
defined over ${\bbbf}_q$ or ${\bbbf}_{q^2}$ is dependent upon the
quadratic character of $-1, d$ and $1-d$, as determined by maps~(\ref{s1}--\ref{s121}).
For example, for $q \equiv 1 \pmod{4}$ one has $\chi_2(-1) = 1$ and so $\sigma_1$ is defined over ${\bbbf}_q$ and
$E_d \sim E_{1-d}$, which was our original observation. We note that the duals
of each of these isogenies are also easily computed.
\subsection{$4$-isogenies of twisted Edwards curves}
One can also map twisted Edwards curves~(\ref{twisted}) to a Legendre form
curve, as given by the following theorem, the proof of which is the same as
the proof of Theorem~\ref{2isogeny}, one having first applied the isomorphism $E_{a,d} \rightarrow
E_{d/a}: (x,y) \mapsto (\sqrt{a}x,y)$.
\begin{theorem} Let $(x,y) \in E_{a,d}$. Then $\psi_{a,d}:E_{a,d} \rightarrow L_{d/a}:$
\[
(x,y) \mapsto \left( \frac{1}{ax^2}, \frac{y(d-a)}{a^{3/2}x(1-y^2)}\right).
\]
The dual of $\psi_{a,d}$ is $\widehat{\psi}_{a,d}:L_{d/a} \rightarrow E_{a,d}:$
\[
(x,y) \mapsto \left( \frac{2\sqrt{a}y}{d-{ax}^2}, \frac{{ay}^2 - {x}^2(a-d)}{{ay}^2 + {x}^2(a-d)}\right).
\]
\end{theorem}
One therefore obtains a set of $4$-isogenies from the isomorphisms of
$L_{d/a}$, exactly as before.
\section{Isomorphisms from $L_d$ to Edwards curves}
In addition to the above $2$-isogeny between $E_d$ and $L_d$, one can also
consider when $L_d$ is birationally equivalent to an Edwards curve, i.e., is
isomorphic to an Edwards curve. Such isomorphisms have two immediate
consequences. Firstly, for each such isomorphism one obtains a $2$-isogeny of
$E_d$ to another Edwards curve $E_{d'}$ via the composition of $\psi_d$ and
the isomorphism, see \S\ref{2isog}. Secondly, one is able to deduce the set of Edwards curves
isomorphic to $E_d$, see \S\ref{Eiso}.
\subsection{Isomorphisms from $L_d$ to $E_{\bar{d}}$}\label{2isog}
Since $L_d:y^2 = x^3 - (1+d)x^2 + dx$, one can transform $L_d$ to the Montgomery
curve
$$M_{A,B}: By^2=x^3+Ax^2+x$$
with $A= - (1+d)/\sqrt{d}$, $B = 1/d\sqrt{d}$ via
$(x,y) \mapsto (x/\sqrt{d},y)$. Using Theorem 3.2 of~\cite{twisted} one then
obtains
\[
\left( \frac{x}{y\sqrt{d}}, \frac{x-\sqrt{d}}{x+\sqrt{d}} \right) \in
E_{-d(1-\sqrt{d})^2,-d(1+\sqrt{d})^2},
\]
which is isomorphic to $E_{\bar{d}}$ with $\bar{d} = \left( \frac{1 + \sqrt{d}}{1 - \sqrt{d}}\right)^2$
with
\[
\rho_d : L_d \rightarrow E_{\bar{d}} : (x,y) \mapsto \left( \sqrt{-1}(1-\sqrt{d})\frac{x}{y}, \frac{x-\sqrt{d}}{x+\sqrt{d}} \right).
\]
Taking the negative root of $d$ in the above transformations gives a second isomorphism, which together we
write as
\[
\rho_{d,\pm} : L_d \rightarrow E_{\bar{d}^{\pm 1}} : (x,y) \mapsto \left(
\sqrt{-1}(1 \mp \sqrt{d})\frac{x}{y}, \frac{x \mp \sqrt{d}}{x \pm \sqrt{d}} \right).
\]
We also have
\[
\widehat{\rho}_{d,\pm}: E_{\bar{d}^{\pm 1}} \rightarrow L_d : (x,y) \mapsto
\left( \pm \sqrt{d}\frac{1+y}{1-y},
\pm \sqrt{-1}\sqrt{d}(1 \mp \sqrt{d})\frac{1+y}{x(1-y)} \right).
\]
Clearly these isomorphisms are only defined over the ground field
if both $-1$ and $d$ are quadratic residues.
Observe that the value $\bar{d}$ is invariant under the substitution $d
\leftarrow 1/d$, hence the $L_d$-isomorphic curve $L_{1/d}$ maps to
$E_{\bar{d}}$ also, but with the $\pm$ isogenies defined instead by
\[
\rho_{1/d,\pm} : L_{1/d} \rightarrow E_{\bar{d}^{\pm 1}} : (x,y) \mapsto
\left( \sqrt{-1}(1 \mp \sqrt{1/d})\frac{x}{y}, \frac{x \mp \sqrt{1/d}}{x \pm \sqrt{1/d}} \right),
\]
with inverse $\widehat{\rho}_{1/d,\pm}: E_{\bar{d}^{\pm 1}} \rightarrow L_{1/d}:$
\[
(x,y) \mapsto \left( \pm \sqrt{1/d}\frac{1+y}{1-y},
\pm \sqrt{-1}\sqrt{1/d}(1 \mp \sqrt{1/d})\frac{1+y}{x(1-y)} \right).
\]
Similarly, one can first map $L_d$ to $L_{\sigma(d)}$ for any $\sigma \in H$, and then apply $\rho_{d,\pm}$ but with the substitution
$d \leftarrow \sigma(d)$ to give $\theta_{\sigma(d),\pm}: L_{d} \rightarrow
L_{\sigma(d)} \rightarrow E_{\overline{\sigma(d)}^{\pm 1}}$.
We thus have twelve isomorphisms $\theta_{\sigma(d),\pm}$ from $L_d$ to the six curves
$E_{\bar{d}_{i}^{\pm 1}}$ for $i \in \{1,2,3\}$, with:
\[
\bar{d}_{1}^{\pm 1} = \left( \frac{1 \pm \sqrt{d}}{1 \mp \sqrt{d}}\right)^2,
\bar{d}_{2}^{\pm 1} = \left( \frac{1 \pm \sqrt{1-d}}{1 \mp \sqrt{1-d}}\right)^2 \ \text{and} \
\bar{d}_{3}^{\pm 1} = \left( \frac{1 \pm \sqrt{\frac{d}{d-1}}}{1 \mp \sqrt{\frac{d}{d-1}}}\right)^2.
\]
As noted above the twelve isomorphisms have only the six image
curves $E_{\bar{d}_{1}^{\pm 1}}, E_{\bar{d}_{2}^{\pm 1}}$ and
$E_{\bar{d}_{3}^{\pm 1}}$, since $d$ and $1/d$ map to $\bar{d}_1$, $1-d$ and
$1/(1-d)$ map to $\bar{d}_2$, and $d/(d-1)$ and $1-1/d$ map to $\bar{d}_3$.
These curves are therefore isomorphic and each has $j$-invariant
\[
\frac{2^8(d^2 - d + 1)^3}{(d(d-1))^2},
\]
which is the Legendre curve $j$-invariant $j_L(d)$.
Taking the composition of $\psi_d$ and an isomorphism from each of the six
pairs of isomorphisms above --- one
from each pair that have the same image --- one obtains
$2$-isogenies of $E_d$ to $E_{\bar{d}_{1}^{\pm 1}}, E_{\bar{d}_{2}^{\pm 1}}$ and
$E_{\bar{d}_{3}^{\pm 1}}$, again defined over ${\bbbf}_{q}$ or ${\bbbf}_{q^2}$ depending
on the quadratic charcter of $-1,d$ and $1-d$,
which we summarise in Theorem~\ref{2isogenies}.
We note that Moody and Shumow have independently given equivalent isogenies~\cite{moody}, having
obtained them using a different approach.
\begin{theorem}\label{2isogenies}
There exist $2$-isogenies of $E_d$ to $E_{\bar{d}_{1}^{\pm 1}}, E_{\bar{d}_{2}^{\pm 1}}$ and
$E_{\bar{d}_{3}^{\pm 1}}$, given by the following maps, respectively:
\begin{itemize}
\item[(a)] $\epsilon_{\bar{d}_{1},\pm} : E_d \rightarrow
E_{\bar{d}_{1}^{\pm 1}}: (x,y) \mapsto \left( \frac{\sqrt{-1}(1 \mp \sqrt{d})}{d-1}
\frac{1-y^2}{xy}, \frac{1 \mp \sqrt{d}x^2}{1 \pm \sqrt{d}x^2} \right)$,
\item[(b)] $\epsilon_{\bar{d}_{2},\pm} : E_d \rightarrow
E_{\bar{d}_{2}^{\pm 1}}: (x,y) \mapsto \left( (1 \mp \sqrt{1-d})xy,
\frac{1-(1 \mp \sqrt{1-d})x^2}{1-(1 \pm \sqrt{1-d})x^2} \right)$,
\item[(c)] $\epsilon_{\bar{d}_{3},\pm} : E_d \rightarrow
E_{\bar{d}_{3}^{\pm 1}}: (x,y) \mapsto \left( \frac{\sqrt{d-1} \mp
\sqrt{d}}{1-d}\frac{x}{y}, \frac{1-\big(d \pm (1-d)\sqrt{\frac{d}{d-1}}\big)x^2}{1-\big(d \mp (1-d)\sqrt{\frac{d}{d-1}}\big)x^2}\right)$.
\end{itemize}
\end{theorem}
Theorem~\ref{2isogenies} allows one to write down the set of $4$-isogenies between $E_d$ and any
$E_{\sigma(d)}$ via isogenies and isomorphisms of Edwards curves only: first
map $E_d \rightarrow E_{\bar{d}_{i}^{\pm 1}}$; second apply an
isomorphism to the relevant $E_{\bar{d}_{j}^{\pm 1}}$; and third use a dual isogeny
to map to $E_{\sigma(d)}$. However, since the Edwards $2$-isogenies implicitly depend
on the $2$-isogeny to $L_d$, the initial derivation given is perhaps the most
natural way to view these $4$-isogenies.
\subsection{Isomorphisms of $E_d$}\label{Eiso}
It is clear from \S\ref{2isog} that the $E_{\bar{d}_{i}^{\pm 1}}$ curves
inherit isomorphisms from the isomorphisms of $L_d$, whereas $E_d$ inherits
isogenies from the isomorphisms of $L_d$
--- in both instances $L_d$ plays a fundamental role. A natural question is whether or not it is possible to exploit the isomorphisms between
$E_{\bar{d}_{i}^{\pm 1}}$ to give the set of curves isomorphic to $E_d$?
Since the $j$-invariant of $E_d$ is~\cite{twisted}
\[
j_E(d) = \frac{16(d^2 + 14d +1)^3}{d(d-1)^4},
\]
it would not seem obvious how to determine the set of isomorphic curves of
$E_d$ from those of $L_d$. However, one can argue as follows.
As above let $\delta = \bar{d}_1(d) = \left( \frac{1 + \sqrt{d}}{1 -
\sqrt{d}}\right)^2$, with $\bar{d}_1$ considered as a function of
$d$. Observe that $d = (\bar{d}_{1}(\delta))^{-1}$, and hence
\[
E_d = E_{\left( \frac{1-\sqrt{\delta}}{1+\sqrt{\delta}} \right)^2}.
\]
Since the curve on the right-hand-side is isomorphic to $E_{\bar{d}_{1}^{\pm 1}(\delta)}$,
$E_{\bar{d}_{2}^{\pm 1}(\delta)}$ and $E_{\bar{d}_{3}^{\pm 1}(\delta)}$, so is
$E_d$. Writing these expressions out in full gives the following theorem.
\begin{theorem}\label{jinv}
Let $E_d$ and $E_{d'}$ be two Edwards curves. Then $E_{d} \cong E_{d'}$ if and
only if
\[
d' \in \Bigg\{d,1/d, \left(\frac{1 \pm d^{1/4}}{1 \mp d^{1/4}}\right)^4,
\left(\frac{1 \pm \sqrt{-1}d^{1/4}}{1 \mp \sqrt{-1}d^{1/4}}\right)^4
\Bigg\}.
\]
\end{theorem}
These six values are naturally implied by Proposition 6.1 of Edwards original exposition~\cite{edwards}.
In particular curve (1) is isomorphic to curve (2) via the map $(x,y) \mapsto (ax,ay)$, with $d = a^4$.
Taking the fourth power of each of the $24$ values given in Edwards' proposition gives the six values listed in Theorem~\ref{jinv}.
It is however interesting that these values can be determined from the isomorphisms of $L_d$ alone.
The above manipulations also show that $E_d \cong L_{\delta}$, via
\[
(x,y) \mapsto \left( \frac{\sqrt{d} + 1}{\sqrt{d} - 1}\cdot \frac{1+y}{1-y},
\frac{2\sqrt{-1}(1+\sqrt{d})}{(1-\sqrt{d})^2}\cdot\frac{1+y}{x(1-y)} \right).
\]
Note that the existence of such an isomorphism is implied by the fact that $j_L(\delta) = j_E(d)$.
\section{The number of isogeny classes of Edwards curves over finite fields}
In this section we derive some results about Edwards
curves~(\ref{edwards}), from results known for the Legendre family of elliptic
curves, which is well-studied. Having established the isogeny between $E_d$ and $L_d$ in
Theorem~\ref{2isogeny}, the validity of this approach is immediate.
In particular we determine the number of isogeny classes of Edwards
curve over the finite field ${\FF_q}$, and in the course of doing so also detail
the number of supersingular curves $E_d({\FF_q})$.
\iffalse
Fortunately, the question of determining
the number of the isogeny classes of Legendre curves over ${\FF_q}$ has been
considered previously and may be deduced by combining results
from~\cite{Katz-ratio} and~\cite{Auer-Top}.
\fi
For the Legendre curve $L_d({\FF_q})$,
we denote the trace of the Frobenius endomorphism
\begin{equation}
-\sum_{x\in{\bbbf}_q}\eta(x(x-1)(x-d))
\end{equation}
by $A(d,{\FF_q})$. Then Equation~(\ref{Legendre-order}) implies
\begin{equation}\label{Leg-order-2}
\#L_d({\FF_q})=q+1-A(d,{\FF_q}),
\end{equation}
and by the Hasse-Weil bound we have
$$
\left|A(d,{\FF_q})\right|\le 2\sqrt{q}.
$$
Thus the number of isogeny classes of the Legendre family of elliptic curves
is the same as the number of integer values of $A$ with
$\left|A\right|\le 2\sqrt{q}$
for which there is a $d$ such that $A(d,{\FF_q})=A$. The following two lemmata
give a satisfactory answer to this question. The first addresses the number
of ordinary isogeny classes and the second addresses the supersingular isogeny classes.
\begin{lemma}\cite{Katz-ratio}\label{ordinary}
Let ${\FF_q}$ be a finite field of odd characteristic, and let $A\in{\bbbz}$
be an integer prime to $p$
(the characteristic of ${\FF_q}$) with $\left|A\right|\le 2\sqrt{q}$.
If $A\equiv q+1 \pmod 4$, then there exists $d\in{\FF_q}\backslash\{0,1\}$
with $A(d,{\FF_q})=A$.
\end{lemma}
\begin{lemma}\cite{Katz-ratio}\label{supersingular}
Let $p$ be an odd prime. Then we have the following assertions.
\begin{itemize}
\item[(i)] If $q=p^{2k+1}$, and $L_d({\FF_q})$ is
supersingular, then $A(d,{\FF_q})=0$.
\item[(ii)] If $q=p^{2k}$, and $L_d({\FF_q})$ is
supersingular, then $A(d,{\FF_q})=\epsilon 2p^k$, where $\epsilon=\pm 1$ is
the choice of sign for which $\epsilon p^k\equiv 1 \pmod 4$.
\end{itemize}
\end{lemma}
Following Katz, we say that each $A$ satisfying the conditions of
Lemma~\ref{ordinary} is {\em unobstructed}, for $q$. From the two lemmata above, the following is immediate.
\begin{cor}
If $q=p^{2k+1}$ and $p\equiv 1 \pmod 4$, then the number of isogeny classes
of Edwards curves over ${\FF_q}$ is
$$
2\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor+2}{4}\right\rfloor-
2\left\lfloor\frac{\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor}{p}\right\rfloor+2}{4}\right\rfloor.
$$
\end{cor}
\begin{proof}
The claim will follow if we prove that there is no supersingular Legendre curve
in this case. Observe that $\#L_d({\FF_q})$ is always divisible by 4, and if $q=p^{2k+1}$, $p\equiv 1 \pmod 4$
and $L_d({\FF_q})$ is supersingular, then from
Lemma~\ref{supersingular}(i) and~(\ref{Leg-order-2}) it follows that $\#L_d\equiv 2\pmod 4$, which is impossible.
\end{proof}
In order to obtain the number of isogeny classes of Edwards curves in the remaining cases
we need to know how the supersingular Legendre curve parameters are
distributed amongst extensions of the prime subfield ${\bbbf}_p$ of ${\bbbf}_q$; again,
there is already a complete answer to this question in the literature. On the one hand, it is
well known that $L_d({\FF_q})$ is a supersinular curve if and only if $d$ is a root of
the Hasse-Deuring polynomial
\[
H_p(x)=(-1)^{\frac{p-1}{2}}\sum_{i=0}^{\frac{p-1}{2}}{{(p-1)/2}\choose{i}}^2x^i,
\]
and on the other hand it is well known that all the roots of Deuring polynomial are in ${\bbbf}_{p^2}$
(see for example \cite[Proposition~2.2]{Auer-Top}).
Using Theorem~\ref{2isogeny} and \cite[Proposition~3.2]{Auer-Top} the following is immediate.
\begin{theorem}\label{supersingular-p}
The number $S_p$ of ${\bbbf}_p$-rational roots of the Deuring polynomial, or equivalently the number
of supersingular Edwards curves over ${\bbbf}_p$, satisfies
\begin{itemize}
\item[(i)] $S_p=0$ if and only if $p\equiv 1 \pmod 4$.
\item[(ii)] $S_3=1$.
\item[(iii)] If $p\equiv 3\pmod 4$ and $p> 3$, then $S_p=3h(-p)$, where $h(-p)$ is the
class number of ${\mathbb Q}(\sqrt{-p})$.
\end{itemize}
\end{theorem}
\begin{cor}
If $p\equiv 3 \pmod 4$ and $q=p^{2k+1}$, then the number of isogeny classes
of Edwards curves over ${\FF_q}$ is
$$
2\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor}{4}\right\rfloor-
2\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor}{4p}\right\rfloor+1.
$$
\end{cor}
\begin{proof}
From Lemma~\ref{supersingular} and Theorem~\ref{supersingular-p} it follows that
there is a single isogeny class of supersingular Legendre curves in this case.
\end{proof}
Similarly we have:
\begin{cor}
If $q=p^{2k}$ for an odd prime $p$, then the number of isogeny classes
of Edwards curves over ${\FF_q}$ is
$$
2\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor+2}{4}\right\rfloor-
2\left\lfloor\frac{\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor}{p}\right\rfloor+2}{4}\right\rfloor+1.
$$
\end{cor}
\begin{proof}
From the fact that all the roots of Hasse-Deuring polynomial are in ${\FF_{p^2}}$
and from Lemma~\ref{supersingular}, it follows that there is
a single isogeny class of supersingular Legendre curves in this case.
\end{proof}
\iffalse
\begin{cor}
If $q=p^{2k}$ and $p\equiv 3 \pmod 4$, then the number of isogeny classes
of Edwards curves over ${\FF_q}$ is
$$
2\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor+2}{4}\right\rfloor-
2\left\lfloor\frac{\left\lfloor\frac{\lfloor 2\sqrt{q}\rfloor}{p}\right\rfloor+2}{4}\right\rfloor+1.
$$
\end{cor}
\begin{proof}
From Lemma~\ref{supersingular} and Theorem~\ref{supersingular-p} it follows that
there is a single isogeny class of supersingular Legendre curves in this case.
\end{proof}
\fi
\section{Isogeny classes of {\em complete} Edwards Curves}\label{complete}
Bernstein and Lange proved that the Edwards addition law is complete, i.e., is well-defined
on all inputs, if and only if $\chi_2(d) = -1$~\cite{bl}.
A natural question to consider is whether there exists a complete Edwards curve
in every isogeny class. In this section we answer this question affirmatively,
relating the number of non-square $d \in {\bbbf}_q \setminus \{0,1\}$ in each
isogeny class to the total number of $d$ in each isogeny class.
\subsection{Katz's ratio results}
While investigating the Lang-Trotter conjecture~\cite{LT}, Katz discovered some remarkable relationships between
the number of $d \in {\bbbf}_q \setminus \{0,1\}$ such that $A(d,{\bbbf}_q) = q + 1 -
\#L_d = A$ for any unobstructed $A$, and the number of $d \in {\bbbf}_q \setminus
\{0,1\}$ such that $A(d,{\bbbf}_q) = -A$~\cite{Katz-ratio}.
In particular, let $N(A) = \#\{ d \in {\bbbf}_{q} \setminus \{0,1\} | A(d,{\bbbf}_q) = A\}$.
Katz proved that for $q \equiv -1 \pmod{4}$, one has
$N(A) = N(-A)$. For $q \equiv 1 \pmod{4}$, this is no longer the case.
Since $A \equiv 2 \pmod 4$, exactly one of $A,-A$ has $q+1 -A \equiv 0 \pmod 8$ --- call it $A$ --- with $q+1+A \equiv 4 \pmod{8}$.
Then $N(A) > N(-A)$. Furthermore, for $q \equiv 5 \pmod{8}$ the ratio
$r = N(A)/N(-A)$ is always one of the integers $2,3$, or $5$,
depending only on the power of $2$ dividing $q+1-A$, as given in:
\begin{theorem}\cite[Theorem 2.8]{Katz-ratio}\label{katz2.8}
Suppose $q \equiv 5 \pmod{8}$. Then
\begin{eqnarray}
\nonumber ord_2(q+1-A) = 3 \Longrightarrow r = 2,\\
\nonumber ord_2(q+1-A) = 4 \Longrightarrow r = 3,\\
\nonumber ord_2(q+1-A) \ge 5 \Longrightarrow r = 5.
\end{eqnarray}
\end{theorem}
For $q \equiv 1 \pmod{8}$ the situation is more complicated. If
$ord_2(q+1-A)=3$ then $r = 2$ as before. Let $\Delta = A^2 - 4q$. For the remaining cases we have:
\begin{theorem}\cite[Theorem 2.11]{Katz-ratio}\label{katz2.11}
Suppose $q \equiv 1 \pmod{8}$, and that $ord_2(q+1-A) \ge 4$. Then
$ord_2(\Delta) \ge 6$, and we have the following results.
\begin{itemize}
\item[(1)] Suppose $ord_2(\Delta) = 2k+1, k \ge 3$. Then $r = 5 - 3/2^{k-2}$.
\item[(2)] Suppose $ord_2(\Delta) = 2k, k \ge 3$. Then
\begin{itemize}
\item[(a)] if $\Delta/2^{2k} \equiv 1 \bmod{8}$, then $r = 5$,
\item[(b)] if $\Delta/2^{2k} \equiv 3 \ \text{or} \ 7 \bmod{8}$, then $r = 5 - 3/2^{k-2}$,
\item[(a)] if $\Delta/2^{2k} \equiv 5 \bmod{8}$, then $r = 5 - 1/2^{k-3}$.
\end{itemize}
\end{itemize}
\end{theorem}
To explain these phenomena, Katz uses the fact that $L_d$ is $2$-isogenous to
the elliptic curve $y^2 = (x+t)(x^2 + x + t)$, $t \ne 0,1/4$, having a point $(0,t)$ of order
$4$ and where $t=(1-d)/4$. Over the $t$-line, this family of
curves with its point $(0,t)$ is the universal curve given with a point of order
$4$. Using this property Katz derives a Deuring-style class number formula
to express the number of $t \in {\bbbf}_q$ such that
$A(t,{\bbbf}_q) = A$. Expressing the same for $-A$ and then computing the
ratio $N(A)/N(-A)$ happens to be far simpler than computing the exact numbers themselves,
as it obviates the need to perform any class group order computations.
However, in the proof no consideration was given (nor was it needed) of the quadratic
character of elements $t$ in a given $N(A)$. Furthermore, since under this
$2$-isogeny we have $t = (1-d)/4$, determining how the corresponding square
and non-square $d$ are distributed between the numerator and denominator of
$N(A)/N(-A)$ is certainly not immediate.
However, we observed (empirically - and then proved) that the following holds. Let $N_{2}(A)$
and $N_{n2}(A)$ be the partition of $N(A)$ into square and non-square $d$
respectively, and similarly for $-A$. For $q \equiv 1 \pmod{4}$, we have
$N_{n2}(A) = N_{n2}(-A) = N(-A)$, i.e., the smallest of the two values
$N(A),N(-A)$. Hence the excess of $N(A)$
over $N(-A)$ consists entirely of square $d$.
For $q \equiv -1 \pmod{4}$ we have
\[
N_{n2}(A) = \begin{cases} N(A) \hspace{7mm} \text{if} \ q+1-A \equiv 4 \pmod{8} \\
N(A)/3 \hspace{3mm} \text{if} \ q+1-A \equiv 0 \pmod{8}. \end{cases}
\]
Since $q \equiv -1 \pmod 4$ we have $N_{n2}(A) = N_{n2}(-A)$ in this case
also. Our proof of these facts is elementary.
\subsection{Proof of claims}
We use the following three lemmata, the
first of which can be found in~\cite[Theorem 8.14]{washington} (see also~\cite[X, Sect. 1]{silverman}):
\begin{lemma}[2-descent]\label{2descent}
Assume $\text{char}({\bbbf}_q) > 2$, and let $E({\bbbf}_q)$ be given by $y^2 = (x -
\alpha)(x-\beta)(x-\gamma)$ with $\alpha,\beta,\gamma \in {\bbbf}_q$, $\alpha \ne
\beta \ne \gamma \ne \alpha$. The map
\[
\phi: E({\bbbf}_q) \longrightarrow {\bbbf}_{q}^{\times}/({\bbbf}_{q}^{\times})^2 \times {\bbbf}_{q}^{\times}/({\bbbf}_{q}^{\times})^2 \times {\bbbf}_{q}^{\times}/({\bbbf}_{q}^{\times})^2
\]
defined by
\begin{eqnarray}
\nonumber (x,y) &\mapsto& (x-\alpha, x-\beta, x- \gamma) \ \text{when} \ y\neq 0\\
\nonumber \mathcal{O} &\mapsto& (1,1,1)\\
\nonumber (e_1,0) &\mapsto& ((e_1 - e_2)(e_1 - e_3), e_1 - e_2, e_1 - e_3)\\
\nonumber (e_2,0) &\mapsto& (e_2 - e_1, (e_2 - e_1)(e_2 - e_3), e_2 - e_3)\\
\nonumber (e_3,0) &\mapsto& (e_3 - e_1, e_3 - e_2, (e_3 - e_1)(e_3 - e_2))
\end{eqnarray}
is a homomorphism, with kernel $2E({\bbbf}_q)$.
\end{lemma}
Applying Lemma~\ref{2descent} to the $2$-torsion points
$(0,0),(1,0)$ and $(d,0)$ of $L_d({\bbbf}_q)$, one can compute the possible
$4$-torsion groups $L_d({\bbbf}_q)[4]$, which depend only on $\chi_2(-1),\chi_2(d)$ and
$\chi_2(1-d)$, giving the following result.
\begin{lemma}\label{4torsion}
For $q \equiv \pm 1 \bmod{4}$, the possible $4$-torsion
groups $L_d({\bbbf}_q)[4]$, are those detailed in Tables 1 and 2 respectively.
\begin{table}[here]
\caption{$q \equiv 1 \bmod{4}$}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline
$\chi_2(d)$ & $\chi_2(1-d)$ & $(L_d({\bbbf}_q)[2] \cap 2L_d({\bbbf}_q))\setminus\{\mathcal{O}\}$ & $L_d({\bbbf}_q)[4]$\\
\hline
$1$ & $1$ & $(0,0),(1,0),(d,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/4{\bbbz}$ \\
$-1$ & $1$ & $(1,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
$1$ & $-1$ & $(0,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
$-1$ & $-1$ & & ${\bbbz}/2{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
\hline
\end{tabular}
\end{center}
\end{table}
\begin{table}[here]
\caption{$q \equiv -1 \bmod{4}$}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline
$\chi_2(d)$ & $\chi_2(1-d)$ & $(L_d({\bbbf}_q)[2] \cap 2L_d({\bbbf}_q))\setminus\{\mathcal{O}\}$ & $L_d({\bbbf}_q)[4]$\\
\hline
$1$ & $1$ & $(1,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
$-1$ & $1$ & $(1,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
$1$ & $-1$ & $(d,0)$ & ${\bbbz}/4{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
$-1$ & $-1$ & & ${\bbbz}/2{\bbbz} \times {\bbbz}/2{\bbbz}$ \\
\hline
\end{tabular}
\end{center}
\end{table}
\end{lemma}
We also use the following easy result, the first part of which was also used by
Katz~\cite[Lemma 2.3]{Katz-ratio}.
\begin{lemma}\label{easyisos}
For $d \in {\bbbf}_q \setminus \{0,1\}$ we have:
\begin{itemize}
\item[(i)] $A(d,{\bbbf}_q) = \chi_2(-1) \cdot A(1-d,{\bbbf}_q)$,
\item[(ii)] $A(d,{\bbbf}_q) = \chi_2(d) \cdot A(1/d,{\bbbf}_q)$.
\end{itemize}
\end{lemma}
\begin{proof}
These are immediate consequences of isomorphisms~(\ref{s1}) and~(\ref{s2}).
\end{proof}
We are now ready to prove our observations.
\begin{theorem}\label{1mod4nonsquare}
For $q \equiv 1 \pmod{4}$, let $A$ be such that $q+1 - A \equiv 0
\pmod{8}$ (and so $q+1 + A \equiv 4 \pmod{8}$). Then $N_{n2}(A) = N_{n2}(-A) =
N(-A)$.
\end{theorem}
\begin{proof}
From Table 1 we see that for any square $d$, $L_d({\bbbf}_q)$ contains a subgroup
of order either $8$ or $16$. As $q + 1 + A \equiv 4 \pmod{8}$, by
Lagrange's theorem we must have $N_{2}(-A) =0$ . Hence all $d$ counted
by $N(-A)$ are necessarily non-square, and since by Lemma~\ref{ordinary} every unobstructed $A$
occurs, we have $N_{n2}(-A) = N(-A)$.
Since ${\bbbf}_{q} \setminus \{0,1,-1\}$ partitions
into a disjoint union of pairs $\{d,1/d\}$,
by Lemma~\ref{easyisos}(ii) for non-square $d$ we have
a bijection between the elements counted by $N_{n2}(-A)$ and those counted by $N_{n2}(A)$, and
hence these numbers are equal.
\end{proof}
\begin{theorem}\label{notsquare2}
For $q \equiv -1 \pmod{4}$, we have
\[
N_{n2}(A) = \begin{cases} N(A) \hspace{7mm} \text{if} \ q+1-A \equiv 4 \pmod{8} \\
N(A)/3 \hspace{3mm} \text{if} \ q+1-A \equiv 0 \pmod{8}. \end{cases}
\]
\end{theorem}
\begin{proof}
We show that the result is true in each isomorphism class.
First, assume $j_L(d) \ne 0,1728$, so that each isomorphism class contains the
six distinct elements in~(\ref{isovalues}). From Table 2 we have that for any
square $d$, $L_d({\bbbf}_q)$ contains a subgroup of order $8$. Hence if $\#L_d({\bbbf}_q) = q + 1 - A \equiv 4
\pmod{8}$, by Lagrange's theorem we must have $N_{2}(A) = 0$. Hence all $d$ counted
by $N(A)$ are non-square, and since every unobstructed $A$
occurs, we have $N_{n2}(A) = N(A)$. This proves the first part of the theorem.
For the second part, we shall show that for each $A$ for which $q+1-A
\equiv 0 \pmod{8}$, square $d$ occur twice as frequently as non-square
$d$ in the counts for both $N(A)$ and $N(-A)$. Abusing notation slightly, when $A(d,{\bbbf}_q) = A$
we write $d \in N(A)$, and simlarly for $N(-A)$.
Let $\#L_d({\bbbf}_q) = q+1-A \equiv 0 \pmod{8}$. Then by Sylow's 1st theorem,
$L_d({\bbbf}_q)$ contains a subgroup of order $8$, and hence $L_d({\bbbf}_q)[8]$
contains at least $8$ points. By Table 2, we can not have
$\chi_2(d)=\chi_2(1-d)=-1$, since $L_d({\bbbf}_q)[4] \cong {\bbbz}_2 \times {\bbbz}_2 = L_d({\bbbf}_q)[2]$ and hence
$|L_d({\bbbf}_q)[2^i]| = 4$ for $i \ge 2$. Hence we have three possibilities for
$(\chi_2(d),\chi_2(1-d))$.
Let $\chi_2(d)=1$ with $d \in N(A)$. Then by Lemma~\ref{easyisos}(ii), $1/d \in
N(A)$ also. By Lemma~\ref{easyisos}(i), $1-d, 1-1/d \in N(-A)$. If
$\chi_2(1-d)=-1$ then by Lemma~\ref{easyisos}(ii) we have $1/(1-d) \in N(A)$, and
$d/(d-1) \in N(-A)$. Hence $\{d,1/d,1/(1-d)\} \in N(A)$ and $\{1-d,1-1/d,d/(d-1)\} \in N(-A)$,
and there are two squares and a non-square in each set, as asserted.
If $\chi_2(1-d)=1$ then by Lemma~\ref{easyisos}(ii) we have instead $1/(1-d) \in
N(-A)$, and $d/(d-1) \in N(A)$. Hence $\{d,1/d,d/(d-1)\} \in N(A)$ and
$\{1-d,1-1/d,1/(1-d)\} \in N(-A)$, and again there are two squares and a non-square in each set.
Finally, if $\chi_2(d)=-1$ and $\chi_2(1-d)=1$, by Lemma~\ref{easyisos} again
we see that if $d \in N(A)$ then $\{d,1-1/d,d/(d-1)\} \in N(A)$ and $\{1/d,1-d,1/(1-d)\} \in N(-A)$.
In all cases $N_{2}(A) = 2N_{n2}(A)$ and $N_{2}(-A) = 2N_{n2}(-A)$, and the second part of the
result follows for these isomorphism classes.
If $j_L(d) = 1728$, i.e., if $d = 2,1/2,-1$, it is easy to see that Lemma~\ref{easyisos}
implies that the trace of Frobenius is zero in all cases. Now $\chi_2(2) = -1$ if $q \equiv
3 \pmod 8$ and is $1$ if $q \equiv 7 \pmod{8}$. In the first case, $q+1 - 0
\equiv 4 \pmod{8}$ and this isomorphism class contributes three elements to
$N_{n2}(0)$ and hence $N(0)$.
In the second case $q+1 - 0 \equiv 0 \pmod{8}$ and this class contributes two
squares and one non-square to $N(0)$.
If $j_L(d) =0$ then $d^2-d+1 =0$, i.e., $d$ and $1/d$ are primitive $6$-th roots
of unity over ${\bbbf}_q$, which are in ${\bbbf}_q$ iff $q \equiv 1 \pmod{6}$. Since $q \equiv -1 \pmod{4}$ we
must have $q \equiv 7 \pmod{12}$. In particular, ${\bbbf}_q$ does not contain any
$12$-th roots of unity and hence $\chi_2(d) = -1$.
Since $1-d = 1/d$, we have $\chi_2(1-d) = \chi_2(1/d) = -1$, and so by Table 2, $L_d({\bbbf}_q)[4] \cong {\bbbz}_2
\times {\bbbz}_2$ and hence $\#L_d({\bbbf}_q) = q+1-A \equiv 4 \pmod{8}$ by the above argument.
By Lemma~\ref{easyisos}(ii), $A(d, {\bbbf}_q) = - A(1/d,{\bbbf}_q)$ and this isomorphism
class contributes one element to $N_{n2}(A)$ and hence $N(A)$, and
one element to $N_{n2}(-A)$ and hence $N(-A)$, whenever this isomorphism class is
defined over ${\bbbf}_q$.
\end{proof}
Since by Lemma~\ref{ordinary} we have $N(A) > 0$ for every unobstructed
integer $A$ for a given $q$, we thus have the following.
\begin{cor}\label{cor}
Let $A$ be an unobstructed integer for $q$. Then there exists at least one quadratic non-residue $d \in {\bbbf}_{q}
\setminus \{0,1\}$ such that $\#E_d({\FF_q}) = q + 1 - A$, and hence there is a
complete Edwards curve in every isogeny class.
\end{cor}
Theorems~\ref{1mod4nonsquare} and~\ref{notsquare2} allow one can compute $N_{n2}(A)$ given $N(A)$, which
can be computed using Katz's Deuring-style class number formula~\cite{Katz-ratio}. In fact for
$q \equiv 1 \pmod{4}$, the formula for $N(-A)$ is far simpler than that for $N(A)$, while for $q \equiv -1 \pmod{4}$,
$N(A)$ and $N_{n2}(A)$ are either equal or differ by a factor of $3$.
To conclude this section, we note that Morain has independently proven the following~\cite[Theorem 17]{morain}.
\begin{theorem}\label{morain}
Let $E({\bbbf}_p): y^2 = x^3 + a_2x^2 + a_4x + a_6$ have three ${\bbbf}_p$-rational
$2$-torsion points. Then there exists a curve $E'({\bbbf}_p)$ isogenous to $E({\bbbf}_p)$ that
is birationally equivalent to a complete Edwards curve.
\end{theorem}
Therefore, if such a curve $E({\bbbf}_q)$ exists in every isogeny class
whose group order is necessarily divisible by $4 = |E({\bbbf}_q)[2]|$, then
Theorem~\ref{morain} implies Corollary~\ref{cor}; Theorem~\ref{mainisog}
provides the missing condition. Furthermore, Morain's proof is constructive, in that from such a curve $E$ one can
explicitly compute a set of isomorphism classes of complete Edwards curves,
based on the structure of the volcano of $2$-isogenies of $E$.
\section{Isogeny classes of {\em original} Edwards curves}
As stated in \S\ref{Eiso}, curves in Edwards' original normal form~(\ref{original}) are
isomorphic to the Bernstein-Lange form~(\ref{edwards}) via $(x,y) \mapsto
(ax,ay)$, with $d = a^4$. Two natural questions to consider are whether or not there exists an original
Edwards curve in every isogeny class, and more specifically how are the original Edwards curves
distributed amongst the isogeny classes? In this section we present answers to
both these questions.
We begin with some definitions. For any unobstructed $A$ for $q$, let $N_{4}(A)$
and $N_{2n4}(A)$ be the number of $d \in N(A)$ that are fourth powers, and
squares but not fourth powers, respectively. For any such $A$ we thus have
\begin{equation}\label{partition}
N(A) = N_{n2}(A) + N_{2n4}(A) + N_4(A).
\end{equation}
Furthermore let $\chi_4(\cdot)$ denote a primitive biquadratic character of ${\bbbf}_q$,
so that $\chi_4(d) = 1$ if and only if there exists an $a \in {\bbbf}_q$ such that
$d = a^4$.
\subsection{Determining $L_d({\bbbf}_q)[8]$}
In the ensuing treatment, we will need to know the possible $8$-torsion subgroups of $L_d({\bbbf}_q)$.
The structure of the $4$-torsion was determined by analysing the halvability
of the $2$-torsion points, using Lemma~\ref{2descent}. Similarly, one can
apply Lemma~\ref{2descent} to the elements of $L_d({\bbbf}_q)[4] \setminus
L_d({\bbbf}_q)[2]$ to determine the structure of the $8$-torsion.
Over the algebraic closure of ${\bbbf}_q$ there are twelve points of order
four; two for each of the three $2$-torsion points $(0,0),(1,0)$ and $(d,0)$:
\begin{eqnarray*}
P_{(0,0),\pm} &=& (\pm \sqrt{d},\sqrt{-1} \sqrt{d}(1 \mp \sqrt{d})),\\
P_{(1,0),\pm} &=& (1 \pm \sqrt{1-d},\sqrt{1-d}(1 \pm \sqrt{1-d})),\\
P_{(d,0),\pm} &=& (d \pm \sqrt{d(d-1)}, \sqrt{d(d-1)}(\sqrt{d} \pm \sqrt{d-1})),
\end{eqnarray*}
along with their negatives (note that one can also prove Lemma~\ref{4torsion}
using these expressions). Applying Lemma~\ref{2descent} to these points gives:
\begin{lemma}\label{8torsion}
The following conditions are both necessary and sufficient for the
points $P_{(0,0),\pm}, P_{(1,0),\pm}$ and $P_{(d,0),\pm}$ respectively, to be halvable:
\begin{itemize}
\item[(i)] $P_{(0,0),\pm} \in 2L_d({\bbbf}_q) \Longleftrightarrow
\pm\sqrt{d},\pm\sqrt{d}-1,\pm \sqrt{d} -d \in ({\bbbf}_{q}^{\times})^2$,
\item[(ii)] $P_{(1,0),\pm} \in 2L_d({\bbbf}_q) \Longleftrightarrow
1 \pm \sqrt{1-d},\pm \sqrt{1-d},1 \pm \sqrt{1-d} -d \in ({\bbbf}_{q}^{\times})^2$,
\item[(iii)] $P_{(d,0),\pm} \in 2L_d({\bbbf}_q) \Longleftrightarrow
d \pm \sqrt{d(d-1)},d \pm \sqrt{d(d-1)}-1,\pm \sqrt{d(d-1)} \in ({\bbbf}_{q}^{\times})^2$.
\end{itemize}
\end{lemma}
\subsection{The case $q \equiv -1 \pmod{4}$}
This is the simplest case, giving rise to the following theorem:
\begin{theorem}\label{pm1main}
If $q \equiv -1 \pmod{4}$, then the following holds:
\begin{itemize}
\item[(i)] Let $a^4 \in {\bbbf}_q \setminus \{0,1\}$. Then $\#L_{a^4}({\bbbf}_q) = p+1-A \equiv 0 \pmod{8}$.
\item[(ii)] Conversely, if $q+1-A \equiv 0 \pmod{8}$ then there exists $a^4 \in
{\bbbf}_q \setminus \{0,1\}$ such that $\#L_{a^4}({\bbbf}_q) = q+1-A$.
\item[(iii)] If $q+1 - A \equiv 0 \pmod{8}$ then $N_{4}(A) = N_{2}(A) = 2N(A)/3$.
\end{itemize}
\end{theorem}
\begin{proof}
Since $a^4$ is a square, by Table 2 we have
$L_{a^4}({\bbbf}_q)[4] \cong {\bbbz}_4 \times {\bbbz}_2$,
and hence by Lagrange's theorem we have $8 \mid \#L_{a^4}({\bbbf}_q)$. This proves $(i)$.
Now let $A$ be any unobstructed integer satisfying $q+1 - A \equiv 0 \pmod{8}$,
and consider the set of all curves $L_d({\bbbf}_q)$ counted by $N(A)$. By Lemma~\ref{ordinary}
this set is non-empty. By Theorem~\ref{notsquare2} we have $N_{2}(A) =
2N(A)/3$. Furthermore, since $q \equiv -1 \bmod{4}$, the map $x^2 \mapsto x^4$ is an automorphism of the set
of squares in ${\bbbf}_q \setminus \{0,1\}$, and hence
$N_{4}(A) = N_{2}(A)$. This proves $(iii)$ and hence $(ii)$.
\end{proof}
\subsection{The case $q \equiv 1 \pmod{4}$}
We have the following theorem, which is proven in the remainder of this section:
\begin{theorem}\label{p14main}
If $q \equiv 1 \pmod{4}$, then the following holds:
\begin{itemize}
\item[(i)] Let $a^4 \in {\bbbf}_q \setminus \{0,1\}$. Then $\#L_{a^4}({\bbbf}_q) = q+1-A \equiv 0 \pmod{16}$.
\item[(ii)] Conversely, if $q+1-A \equiv 0 \pmod{16}$ then there exists $a^4 \in
{\bbbf}_q \setminus \{0,1\}$ such that $\#L_{a^4}({\bbbf}_q) = q+1-A$.
\item[(iii)] If $q+1 - A \equiv 0 \pmod{16}$ then $N_{4}(A) = N(A) - 2N(-A)$.
\end{itemize}
\end{theorem}
Note that the implication in $(iii)$ is equivalent to $N_4(A)/N(A) = 1 - 2/r$, where $r$ is Katz's
ratio $N(A)/N(-A)$. Using Theorem~\ref{1mod4nonsquare} and~(\ref{partition}),
this is equivalent to $N_4(A) = N_{n2}(A) + N_{2n4}(A) + N_{4}(A) - 2N_{n2}(A)$, or
\begin{equation}\label{mainbijection}
N_{2n4}(A) = N_{n2}(A).
\end{equation}
Equation~(\ref{mainbijection}) in fact holds for all $A$ such that $q+1 - A \equiv
0 \pmod{8}$, and seems to be non-trivial. We will prove it by constructing a bijection between
the sets of curve parameters of each type. Once this equality is proven, part $(ii)$ follows easily.
The idea behind the proof of Equation~(\ref{mainbijection}) is a natural extension of the bijection-based proofs
of \S\ref{complete} , which used the isomorphisms given in Lemma~\ref{easyisos}. Rather than use
isomorphisms defined over ${\bbbf}_q$, which are isogenies of degree one, we use
isogenies of degree two. In particular we consider the isomorphism classes
of curves arising from two $2$-isogenies of $L_d$: the first being ``divide by the ${\bbbz}/2{\bbbz}$ generated
by $(0,0)$'' when $d \in N_{2n4}(A)$, and the second being ``divide by the ${\bbbz}/2{\bbbz}$ generated
by $(1,0)$'' when $d \in N_{n2}(A)$, which are dual to one another.
We begin with a short proof of part $(i)$.
\vspace{4mm}
\newline
\noindent
{\em Proof of (i):} Let $d = a^4$. Since $\chi_2(d) = 1$, by Table 1, if $\chi_2(1-d)=1$ then
$L_d({\bbbf}_q)[4] \equiv {\bbbz}_4 \times {\bbbz}_4$ and hence $16 \mid \#L_d({\bbbf}_q)$. If
$\chi_2(1-d)=-1$ then by Table 1 neither of $(1,0)$ or $(d,0)$ are
halvable, and we claim that precisely one of $P_{(0,0),\pm}$ is
halvable. As $\chi_2(-1) = 1$, by Lemma~\ref{8torsion}, $P_{(0,0),+}$ is halvable if and only if $\sqrt{d},
\sqrt{d}-1$ are both square, while $P_{(0,0),-}$ is halvable if and only if $-\sqrt{d},
-\sqrt{d}-1$ are both square. Since $\chi_4(d) = 1$, both $\pm \sqrt{d}$ are
square. Furthermore, as $1-d = (1+\sqrt{d})(1-\sqrt{d}) =
(-\sqrt{d}-1)(\sqrt{d}-1)$, precisely one of these factors is square as
$\chi_2(1-d)=-1$ by assumption. This gives rise
to a point of order $8$. Therefore ${\bbbl}_d({\bbbf}_q)[8] \cong {\bbbz}_8 \times {\bbbz}_2$
and hence $16 \mid \#L_d({\bbbf}_q)$ in this case too. This completes the proof of $(i)$.
\vspace{4mm}
\newline
We now exhibit a bijection to prove~(\ref{mainbijection}), assuming $q+1-A
\equiv 0 \pmod{8}$.
\begin{lemma}
Let $A$ satisfy $q+1 - A \equiv 0 \pmod{8}$. Then there exists an injection
from $N_{2n4}(A)$ to $N_{n2}(A)$.
\end{lemma}
\begin{proof}
Note that if $d \in N_{2n4}(A)$ then by Table 1 we necessarily have $q+1-A \equiv 0
\pmod{8}$. Let $\xi_d: L_d \rightarrow L_d/\langle (0,0) \rangle$ and let $E^d = \xi_d(L_d)$.
Using V\'{e}lu's formula~\cite{velu}, $E^d$ has equation $y^2 = x^3 - (d+1)x^2 - 4dx + 4d(d+1) = (x - (d+1))(x -
2\sqrt{d})(x + 2\sqrt{d})$, and
\[
\xi_d(x,y) = ( x + d/x, y(1 - d/x^2)).
\]
In particular, $(1,0),(d,0) \in L_d$ are both mapped to $(d+1,0) \in E^d$, and
hence $L_d$ is isomorphic to $E^d/\langle (d+1,0) \rangle$.
Labelling the abscissae of the order $2$ points of $E^d$ by $e_1= d+1,e_2=2\sqrt{d}$ and
$e_3=-2\sqrt{d}$, one sees~(\cite{washington}) that $E^d$ has six isomorphic Legendre curves, each given by a permutation
of $(e_1,e_2,e_3)$ with paramater $\lambda = (e_3 - e_1)/(e_2 - e_1)$, and
\[
(x,y) \mapsto \bigg( \frac{x-e_1}{e_2-e_1}, \frac{y}{(e_2-e_1)^{3/2}} \bigg).
\]
Each of these isomorphisms is defined over ${\bbbf}_q$ if and only if $\lambda \in {\bbbf}_q$ and
$\chi_2(e_2 - e_1) =1$~\cite{washington}. For $d \in N_{2n4}(A)$, the two $E^d$-isomorphic Legendre curves used in
the bijection are given in Table 3.
\begin{table}[here]
\caption{$L_d/\langle (0,0) \rangle$-isomorphic Legendre curves in $N_{n2}(A)$
for $d \in N_{2n4}(A)$}\label{tab3}
\begin{center}
\begin{tabular}{c|c|c|c}
\hline
$(e_1,e_2,e_3)$ & $\lambda$ & $(e_2-e_1)$ & $\chi_2(e_2-e_1)$ \\
\hline
$(2\sqrt{d},d+1, -2\sqrt{d})$ & $\lambda_1 = \frac{-4\sqrt{d}}{(1-\sqrt{d})^2}$
& $(1-\sqrt{d})^2$ & $1$ \\
$(-2\sqrt{d},d+1,2\sqrt{d})$ & $\lambda_2 =\frac{4\sqrt{d}}{(1+\sqrt{d})^2}$
& $(1+\sqrt{d})^2$ & $1$ \\
\hline
\end{tabular}
\end{center}
\end{table}
Observe that $\lambda_1(d),\lambda_2(d) \in N_{n2}(A)$ since $\chi_4(d) \ne 1$.
Note also that $\lambda_1 = 1 - \delta$, with $\delta$ as given in \S\ref{Eiso}, and hence
this isomorphism class is precisely that of $E_d$; indeed we have $j_L(\delta) = j_E(d)$.
Thus $E^d \cong E_d$, explaining our choice of notation.
Abusing notation slightly, we refer to the isomorphisms
$E^d \rightarrow L_{\lambda_1(d)}$ and $E^d \rightarrow L_{\lambda_2(d)}$ by $\lambda_1(d)$ and $\lambda_2(d)$
respectively. Note that both $\lambda_1(d)$ and $\lambda_2(d)$ map $(d+1,0) \in E^d$ to
$(1,0) \in L_{\lambda_1(d)},L_{\lambda_2(d)}$.
Furthermore, if $d$ is replaced with $1/d$ in Table 3, then each $\lambda_i(d)$
remains invariant. Hence $L_{1/d}$ maps to $\lambda_1(d),\lambda_2(d)$ as well, via
$\xi_{1/d}(L_{1/d}) = E^{1/d}$, and
the point $(1/d + 1,0) \in E^{1/d}$ maps to $(1,0) \in L_{\lambda_1(d)},L_{\lambda_2(d)}$.
As $1/d \in N_{2n4}(A)$, this means we
have a map from the pair $\{d,1/d\} \subset N_{2n4}(A)$ to the pair
$\{\lambda_1(d),\lambda_2(d)\} \subset N_{n2}(A)$.
Note that $d,1/d$ are distinct, unless $d=-1$ and $q \equiv 5 \pmod{8}$, in which case we have
$\lambda_1(-1) = \lambda_2(-1) = 2$ with $\chi_2(2) = -1$ and hence $2 \in
N_{n2}(A)$. So in this exceptional case, we have an injection.
In the general case we thus have two pairs of maps:
\begin{eqnarray}
\nonumber \lambda_1(d) \circ \xi_d : L_d &\longrightarrow& L_{\lambda_1(d)},\\
\nonumber \lambda_2(d) \circ \xi_d : L_d &\longrightarrow& L_{\lambda_2(d)},\\
\nonumber \lambda_1(1/d) \circ \xi_{1/d} : L_{1/d} &\longrightarrow& L_{\lambda_1(1/d)},\\
\nonumber \lambda_2(1/d) \circ \xi_{1/d} : L_{1/d} &\longrightarrow& L_{\lambda_2(1/d)},
\end{eqnarray}
with $L_{\lambda_1(d)} = L_{\lambda_1(1/d)}$ and $L_{\lambda_2(d)} = L_{\lambda_2(1/d)}$.
We claim the above four maps taken together form an injective map from pairs $\{d,1/d\}$ to
pairs $\{\lambda_1(d),\lambda_2(d)\}$. Indeed suppose that for $d' \in N_{2n4}(A)$ we have
\[
\lambda_1(d') =
\lambda_1(d) \ \text{or} \ \lambda_2(d),
\ \text{or} \
\lambda_2(d') =
\lambda_1(d) \ \text{or} \ \lambda_2(d).
\]
Then $\sqrt{d'} = \pm \sqrt{d}$ or $\sqrt{d'} = \pm 1/\sqrt{d}$, i.e., $d' = d$
or $d' = 1/d$, and hence the map is injective on the stated pairs.
\end{proof}
Now consider the reverse direction, which is almost immediate.
\begin{lemma}
Let $A$ satisfy $q+1 - A \equiv 0 \pmod{8}$. Then there exists an injection
from $N_{n2}(A)$ to $N_{2n4}(A)$.
\end{lemma}
\begin{proof}
Let $e \in N_{n2}(A)$. For $q+1-A \equiv 0 \pmod{8}$, by Table 1 we must have $\chi_2(e) = -1$ and
$\chi_2(1-e)=1$. The only isomorphism defined over ${\bbbf}_q$ in this case
maps $ L_e \longrightarrow L_{e/(e-1)}$ (see (15)).
Therefore if $e \in N_{n2}(A)$, then $\frac{e}{e-1} \in N_{n2}(A)$.
Indeed $\lambda_2(d) = \lambda_1(d)/(\lambda_1(d)-1)$ (and $\lambda_1(d) = \lambda_2(d)/(\lambda_2(d)-1)$).
Since $\lambda_1(d)$ and $\lambda_2(d)$ map the $\hat{\xi}_d$-generating element $(d+1,0)$
of $E^d$ to $(1,0)$ in $L_{\lambda_1}$ and $L_{\lambda_2}$ (and similarly
$(1/d+1,0) \in E^{1/d}$ to $(1,0)$), the dual $\hat{\xi}_d$ of $\xi_d$ applied
to the isomorphism class representative $L_{e}$ is given by $L_e/\langle (1,0) \rangle$, and
similarly for $L_{e/(e-1)}$. Hence if $e = \lambda_1(d)$ or $\lambda_2(d)$, then $\hat{\xi}_d$
maps elements of $N_{n2}(A)$ to the original isomorphism class of $L_d$.
We now analyse this map and identify which curves in the resulting isomorphism class are relevant.
For the sake of generality let $\gamma_e : L_e \longrightarrow L_e/\langle (1,0)
\rangle$ and let $F^e = \gamma_e(L_e)$. Using V\'{e}lu's formula $F^e$ has equation
$y^2 = x^3 - (e+1)x^2 - (6e-5)x - 4e^2 +7e -3 =
(x - (e-1))(x -(1 + 2\sqrt{1-e}))(x -(1 - 2\sqrt{1-e}))$, and
\[
\gamma_e(x,y) := \bigg( x + \frac{1-e}{x-1}, y\bigg(1 - \frac{1-e}{(x-1)^2}\bigg) \bigg).
\]
Note that $\gamma_e(0,0) = \gamma_e(e,0) = (e-1,0)$. For $e \in N_{n2}(A)$, the two $F^e$-isomorphic Legendre curves used in
the bijection are given in Table 4.
\begin{table}[here]
\caption{Two $L_e/\langle (1,0) \rangle$-isomorphic Legendre curves in
$N_{2n4}(A)$ for $e \in N_{n2}(A)$ and $q+1-A \equiv 0 \pmod{8}$}
\begin{center}
\begin{tabular}{c|c|c}
\hline
$(e_1,e_2,e_3)$ & $\mu$ & $(e_2-e_1)$ \\
\hline
$(e-1,1 + 2\sqrt{1-e},1-2\sqrt{1-e})$ & $\mu_1 = \big(\frac{1-\sqrt{1-e}}{1+\sqrt{1-e}} \big)^2$
& $(1+\sqrt{1-e})^2$ \\
$(e-1,1-2\sqrt{1-e},1+2\sqrt{1-e})$ & $\mu_2 = \big(\frac{1+\sqrt{1-e}}{1-\sqrt{1-e}} \big)^2$
& $(1 - \sqrt{1-e})^2$ \\
\hline
\end{tabular}
\end{center}
\end{table}
Observe that $\chi_2(e_2 - e_1) = 1$ in each case, and the same is true for $\mu_1(e),\mu_2(e)$.
Furthermore, $\mu_1(e),\mu_2(e)$ are both in
$N_{2n4}(A)$ since $(1 \pm \sqrt{1-e})/(1 \mp \sqrt{1-e})$ is not square.
Indeed, since $1-e$ is square, write $1-e = b^2$ so that $e = 1- b^2 = (1+b)(1-b)$.
Therefore $-1 = \chi_2(e) = \chi_2(1+b)\chi_2(1-b) = \chi_2(1+b)/\chi_2(1-b) = \chi_2((1+b)/(1-b))$.
Again abusing notation slightly, we refer to the isomorphisms
$F^e \rightarrow L_{\mu_1(e)}$ and $F^e \rightarrow L_{\mu_2(e)}$ by $\mu_1(e)$ and $\mu_2(e)$
respectively.
Note that both $\mu_1(e)$ and $\mu_2(e)$ map $(e-1,0) \in F^e$ to
$(0,0) \in L_{\mu_1(e)},L_{\mu_2(e)}$.
Furthermore, if $e$ is replaced with $e/(e-1)$ in Table 4, then each $\mu_i(e)$
remains invariant. Hence $L_{e/(e-1)}$ maps to $\mu_1(e),\mu_2(e)$ as well, via
$\gamma_{e/(e-1)}(L_{e/(e-1)}) = F^{e/(e-1)}$, and
the point $(e/(e-1) - 1,0) \in F^{e/(e-1)}$ maps to $(0,0) \in L_{\mu_1(e)},L_{\mu_2(e)}$.
As $e/(e-1) \in N_{n2}(A)$, this means we
have a map from the pair $\{e,e/(e-1)\} \subset N_{n2}(A)$ to the pair
$\{\mu_1(e),\mu_2(e)\} \subset N_{2n4}(A)$.
Note that $e,e/(e-1)$ are distinct, unless $e = 2$ and $q \equiv 5 \pmod{8}$, in which case we have
$\mu_1(2) = \mu_2(2) = -1$ with $\chi_4(-1) \ne 1$ and hence $-1 \in
N_{2n4}(A)$. So in this exceptional case, we have an injection (in fact the inverse of the previous injection).
In the general case we thus have two pairs of maps:
\begin{eqnarray}
\nonumber \mu_1(e) \circ \gamma_e : L_e &\longrightarrow& L_{\mu_1(e)},\\
\nonumber \mu_2(e) \circ \gamma_e : L_e &\longrightarrow& L_{\mu_2(e)},\\
\nonumber \mu_1(e/(e-1)) \circ \gamma_{e/(e-1)} : L_{e/(e-1)} &\longrightarrow& L_{\mu_1(e/(e-1))},\\
\nonumber \mu_2(e/(e-1)) \circ \gamma_{e/(e-1)} : L_{e/(e-1)} &\longrightarrow& L_{\mu_2(e/(e-1))},
\end{eqnarray}
with $L_{\mu_1(e)} = L_{\mu_1(e/(e-1))}$ and $L_{\mu_2(e)} = L_{\mu_2(e/(e-1))}$.
We claim the above four maps taken together form an injective map from pairs $\{e,e/(e-1)\}$ to
pairs $\{\mu_1(e),\mu_2(e)\}$. Indeed suppose that for $e' \in N_{n2}(A)$ we have
\[
\mu_1(e') =
\mu_1(e) \ \text{or} \ \mu_2(e),
\ \text{or} \
\mu_2(e') =
\mu_1(e) \ \text{or} \ \mu_2(e).
\]
Then $e' = e$ or $e' = e/(e-1)$, and hence the map is injective on the stated pairs.
\end{proof}
We have thus proven:
\begin{theorem}\label{bijproof}
Let $A$ satisfy $q+1 - A \equiv 0 \pmod{8}$. Then there exists a bijection
between $N_{2n4}(A)$ and $N_{n2}(A)$.
\end{theorem}
Furthermore, using the above definitions one can check that
\[
\mu_1(\lambda_1(d)) = \mu_1(\lambda_2(d)) = d, \ \text{and} \ \mu_2(\lambda_1(d)) = \mu_2(\lambda_2(d)) = 1/d,
\]
and similarly
\[
\lambda_1(\mu_1(e)) = \lambda_1(\mu_2(e)) = e/(e-1), \ \text{and} \ \lambda_2(\mu_1(e)) = \lambda_2(\mu_2(e)) = e,
\]
and that
\begin{eqnarray}
\nonumber (\mu_2(\lambda_1(d)) \circ \gamma_{\lambda_1(d)}) \circ (\lambda_1(d) \circ \xi_d) &=& [2]
\ \text{on} \ L_d,\\
\nonumber (\mu_2(\lambda_2(d)) \circ \gamma_{\lambda_2(d)}) \circ (\lambda_2(d) \circ \xi_d) &=& [2]
\ \text{on} \ L_d,\\
\nonumber (\lambda_1(\mu_1(e)) \circ \xi_{\mu_1(e)}) \circ (\mu_1(e) \circ \gamma_{e}) &=& [2]
\ \text{on} \ L_e,\\
\nonumber (\lambda_1(\mu_2(e)) \circ \xi_{\mu_2(e)}) \circ (\mu_2(e) \circ \gamma_{e}) &=& [2]
\ \text{on} \ L_e.
\end{eqnarray}
Observe that if one substitutes $d \in N_{2n4}(A)$ for $e$ in the latter two maps, then one
obtains $2$-isogenies from $L_d$ to $L_{\mu_1(d)}$,$L_{\mu_2(d)}$, however
$\mu_1(d), \mu_2(d) \not\in N_{n2}(A)$, so the bijection can only be used in the manner proven.
So while the bijection principally relies on a $2$-isogeny and its dual, this
alone is insufficient; one needs to also consider the isomorphism class
representatives used, which is natural given that we are considering Legendre
curve parameters $d$ rather than isomorphism classes of curves.
With regard to Theorem~\ref{p14main}, note that Theorem~\ref{bijproof}
directly implies Theorem~\ref{p14main}(iii).
Let $A$ be any unobstructed integer satisfying $q+1 - A \equiv 0 \pmod{16}$, and consider the set of all
curves $L_d({\bbbf}_q)$ counted by $N(A)$. By Lemma~\ref{ordinary} this set is non-empty.
Theorems~\ref{katz2.8} and~\ref{katz2.11} show that for $q+1-A \equiv 0 \pmod{16}$ the ratio $N(A)/N(-A)
> 2$ and thus $N_4(A) = N(A) - 2N(-A) > 0$, which thus proves part $(ii)$ and completes the proof.
\section{Curves defined using a ratio of two quadratics}\label{quad}
Following on from \S\ref{pointcount} where we expressed the equation defining
$E_d$ in the form~(\ref{frac}), in this section we briefly discuss curves defined using a ratio of two
quadratic polynomials or a ratio of a quadratic and a linear polynomial. We
demonstrate that one can derive an addition formula for these types
of curves and prove for them results similar to the results of the preceeding sections.
\subsection{Ratio of two quadratics}
Let $f(x)=a_1x^2+b_1x+c_1, g(x)=a_2x^2+b_2x+c_2\in {\FF_q}[x]$ be as in Lemma~\ref{q-c-s},
$a_1,a_2$ both non-zero, and define a curve by the equation
\begin{equation}\label{curveC}
C/{\bbbf}_q: y^2={\frac{a_1x^2+b_1x+c_1}{a_2x^2+b_2x+c_2}}.
\end{equation}
Notice that writing the curve equation as a ratio of two quadratics is just for the sake of the
exposition and it is understood that $C/{\FF_q}$ is the projective curve defined by the equation
$$
(a_2x^2+b_2xz+c_2z^2)y^2=a_1x^2z^2+b_1xz^3+c_2z^4.
$$
Now suppose that
$$
f(x)=a_1(x-\omega_1)(x-\omega_2),
$$
and
$$
g(x)=a_2(x-\omega_3)(x-\omega_4).
$$
The conditions of Lemma~\ref{q-c-s} imply that $\omega_1,\omega_2,\omega_3,\omega_4$ are
pairwise distinct. This implies that there is a linear fractional transformation
$$
\phi: x \mapsto \frac{u_1x+u_2}{u_3x+u_4}\;\;\;\;u_i\in\overline{{\FF_q}},
$$
which maps $\omega_1,\omega_2,\omega_3,\omega_4$ to $\mu,-\mu,1/\mu,-1/\mu$ provided
that the cross-ratio condition
$$
\frac{(\omega_1-\omega_3)(\omega_2-\omega_4)}{(\omega_2-\omega_3)(\omega_1-\omega_4)}=
\left(\frac{\mu^2-1}{\mu^2+1}\right)^2
$$
is satisfied (see~[Chapter 4]\cite{needham}). The map $\phi$ induces the map
$$
\psi: x \mapsto \frac{-u_4x+u_2}{u_3x-u_1}$$
which in turn induces an isomorphism of the function field $\overline{{\FF_q}}(C)$ and the function field
of the curve $E^\mu$, $\overline{{\FF_q}}(E^\mu)$, where $E^{\mu}$ is defined by:
$$
y^2=\frac{x^2-\mu^2}{x^2-\frac{1}{\mu^2}}.
$$
$E^\mu$ is clearly isomorphic to the original Edwards curve~(\ref{original}). Thus
$\overline{{\FF_q}}(C)$ is an elliptic function field and hence the desingularization
of $C$ yields an elliptic curve. One can obtain results similar to the ones
proven in~\cite{edwards} for the curve $C$. For example, one can obtain an addition formula
for the points on $C$ by using the Edwards curve addition formula and the map $\phi$,
as $\phi$ induces a group homomorphism between the group of points on $C$ and
the group of points on $E^\mu$.
\subsection{Ratio of a quadratic and a linear polynomial}
Now suppose that for the curve~(\ref{curveC}) we have $a_2=0, b_2\neq 0$, giving the
corresponding curve
\begin{equation}\label{quadlin}
C'/{\bbbf}_q: y^2={\frac{a_1x^2+b_1x+c_1}{b_2x+c_2}}.
\end{equation}
Then there is a linear fractional transformation
\[
\varphi: x \mapsto \frac{u_{1}^{'}x+u_{2}^{'}}{u_{3}^{'}x+u_{4}^{'}},\;\; u_{i}^{'}\in \overline{{\bbbf}_q},
\]
which maps $C'$ to a curve of the form~\ref{curveC} defined by a ratio of two quadratics, and
which induces an isomorphism between the function fields of $C'$ and a curve
of the form~\ref{curveC}.
Thus our discussion in the previous section applies to curves defined
using the ratio of a quadratic and a linear polynomial.
\subsubsection{Huff curves}
The Huff's model of elliptic curves introduced by Huff~\cite{huff}
which has recently captured the interest of the cryptographic community~\cite{joy,wu-feng}
can be transformed to one of the form~(\ref{quadlin}).
In particular, the Huff's curve, defined by the equation
\[
H_{a,b}/{\FF_q}: ax(y^2-1)=by(x^2-1),
\]
is transformed to the curve
\[
y^2=\frac{bt^2+at}{at+b},
\]
by setting $xy=t$. Thus one can generate the Huff's curve addition law using the process
outlined in the previous section. Furthermore, whenever a curve family is
${\bbbf}_q$-isomorphic to an Edwards or Legendre curve, one can deduce some
properties of the isogeny classes. For example, we have~\cite{wu-feng}
\[
H_{a,b} \cong E_{\big(\frac{a-b}{a+b}\big)^2} \ \text{over} \ {\bbbf}_q,
\]
and so applying Theorems~\ref{1mod4nonsquare} and~\ref{notsquare2} we conclude that for any
unobstructed $A$, if $q+1 - A \equiv 0 \pmod{8}$ then there exists a Huff's curve over ${\bbbf}_q$ with that cardinality.
One can also apply the results of this paper directly to the Jacobi intersection family~\cite{jacobi}
\[
x^2 + y^2 = 1 \ \text{and} \ dx^2 + z^2 = 1,
\]
since this family has $j$-invariant $j_L(d)$.
\begin{remark}
A new single-parameter family of elliptic curves was introduced
in~\cite{Fre-Cas} (amongst more than $50,000$ others) defined
by the curve equation
\[
Ax+x^2-xy^2+1=0,
\]
which enjoys a uniform $x$-coordinate addition formula.
The curve equation can be rewritten as
\[
y^2=\frac{x^2+Ax+1}{x}.
\]
Hence one can obtain addition formula for this family of curves using the addition law of
Edwards curves, although we do not claim that this method generates the most
efficient group law.
\end{remark}
\section{Concluding remarks}
We have identified the set of isogeny classes of Edwards curves over finite fields of odd characteristic,
and have found the proportion of parameters $d$ in each isogeny class which give rise to complete Edwards curves.
Furthermore, we have identified the set of isogeny classes of original Edwards curves, and proven similar proportion results
for this sub-family of curves.
Although not included in the paper, by analysing the $4$- and $8$-torsion of
Legendre curves, and using variants of the established bijections, we were able to prove parts of
Katz's ratio theorems. We believe an interesting and challenging problem is
whether or not the methods of this paper can be developed to provide an alternative proof for
all parts of Katz's ratio theorems; and conversely, can Katz's methods be used to find relationships between
$N_{2^k}(A)$ and $N(A)$ similar to those proven in
Theorems~\ref{1mod4nonsquare},~\ref{notsquare2},~\ref{pm1main}(iii) and~\ref{p14main}(iii), for $k>2$ and $q \equiv 1 \pmod{2^k}$?
\section*{Acknowledgements}
The authors would like to thank Steven Galbraith for offering some useful initial
pointers, and for answering all our questions.
\bibliographystyle{plain}
|
\section{Introduction}\label{sec:intro}
An $n$-step uniform random walk is a walk in the plane that starts at the origin and
consists of $n$ steps of length $1$ each taken into a uniformly random
direction. The study of such walks largely originated with Pearson more than a
century ago \cite{Pea05,Pea05b,Pea06} who posed the problem of determining the
distribution of the distance from the origin after a certain number of steps.
In this paper, we study the (radial) densities $p_n$ of the distance travelled in $n$
steps. This continues research commenced in \cite{bnsw-rw, bsw-rw2} where the focus
was on the moments of these distributions:
\begin{equation*}
W_n(s):=\int_0^n p_n(t) t^{s} \, \md t.
\end{equation*}
The densities for walks of up to $8$ steps are depicted in Figure \ref{fig:pn}.
As established by Lord Rayleigh \cite{Ray05}, $p_n$ quickly approaches the
probability density $\frac{2 x}{n} e^{-x^2/n}$ for large $n$. This limiting
density is superimposed in Figure \ref{fig:pn} for $n\ge5$.
\begin{figure}[htbp]
\begin{center}
\subfigure[$p_3$\label{fig:p3}]{\includegraphics[width=0.3\textwidth]{p3}}
\hfil
\subfigure[$p_4$\label{fig:p4}]{\includegraphics[width=0.3\textwidth]{p4}}
\hfil
\subfigure[$p_5$\label{fig:p5}]{\includegraphics[width=0.3\textwidth]{p5a}}\\
\subfigure[$p_6$\label{fig:p6}]{\includegraphics[width=0.3\textwidth]{p6a}}
\hfil
\subfigure[$p_7$\label{fig:p7}]{\includegraphics[width=0.3\textwidth]{p7a}}
\hfil
\subfigure[$p_8$\label{fig:p8}]{\includegraphics[width=0.3\textwidth]{p8a}}
\caption{Densities $p_n$ with the limiting behaviour superimposed for $n\ge5$.}
\label{fig:pn}
\end{center}
\end{figure}
Closed forms were only known in the cases $n=2$ and $n=3$. The evaluation, for $0\le x \le 2$,
\begin{equation}\label{eq:p2}
p_2(x) = \frac{2}{\pi \sqrt{4-x^2}}
\end{equation}
is elementary. On the other hand, the density $p_3(x)$ for $0\le x \le 3$ can
be expressed in terms of elliptic integrals by
\begin{equation}\label{eq:p3ell}
p_3(x) = \Re \left(\frac{\sqrt{x}}{\pi^2}\, K\left(\sqrt{\frac{(x+1)^3(3-x)}{16x}}\right)\right),
\end{equation}
see, e.g., \cite{Pea06}. One of the main results of this paper is a closed form
evaluation of $p_4$ as a hypergeometric function given in Theorem
\ref{thm:p4hyp04}. In \eqref{eq:p3hyp} we also provide a single hypergeometric
closed form for $p_3$ which, in contrast to \eqref{eq:p3ell}, is real and valid
on all of $[0,3]$. For convenience, we list these two closed forms here:
\begin{align}
p_3(x) &= \frac{2\sqrt{3}}{\pi}\,\frac{x}{\left(3+x^2\right)}
\pFq21{\frac13,\frac23}{1}{\frac{x^2 \left( 9-x^2 \right)^2}{\left( 3+x^2 \right)^3}}, \\
p_4(x) &= \frac2{\pi^2}\,\frac{\sqrt {16-x^2}}x\,
\Re \pFq32{\frac12, \frac12, \frac12}{\frac56,\frac76}
{\frac {\left( 16-x^2 \right)^3}{108 x^4}}.
\end{align}
We note that while \emph{Maple} handled these well to high
precision, \emph{Mathematica} struggled, especially with the analytic
continuation of the $_3F_2$ when the argument is greater than 1.
A striking feature of the 3- and 4-step random walk densities is
their modularity. It is this circumstance which not only allows us
to express them via hypergeometric series, but also makes them a
remarkable object of mathematical study.
This paper is structured as follows: In
Section \ref{sec:pn} we give general results for the densities $p_n$
and prove for instance that they satisfy certain linear differential
equations. In Sections \ref{sec:p3}, \ref{sec:p4}, and \ref{sec:p5}
we provide special results for $p_3$, $p_4$, and $p_5$ respectively.
Particular interest is taken in the behaviour near the points where
the densities fail to be smooth. In Section \ref{sec:Wndiff} we
study the derivatives of the moment function and make a connection
to multidimensional Mahler measures. Finally in Section \ref{sec:Wn}
we provide some related new evaluations of moments and so resolve a
central case of an earlier conjecture on convolutions of moments in
\cite{bsw-rw2}.
The amazing story of the appearance of
Theorem~\ref{thm:deleadingcoeffx} is worth mentioning here. The
theorem was a conjecture in an earlier version of this manuscript,
and one of the present authors communicated it to D.~Zagier. That
author was surprised to learn that Zagier had already been asked for
a proof of exactly the same identities a little earlier, by
P.~Djakov and B.~Mityagin.
Those authors had in fact proved the theorem already in 2004 (see
\cite[Theorem 4.1]{dm1} and \cite[Theorem 8]{dm2}) during their
study of the asymptotics of the spectral gaps of a Schr\"odinger
operator with a two-term potential --- their proof was indirect, so
that we should never have come across the identities without the
accident of asking the same person the same question! Djakov and
Mityagin asked Zagier about the possibility of a direct proof of
their identities (the subject of Theorem~\ref{thm:deleadingcoeffx}),
and he gave a very neat and purely combinatorial answer. It is this
proof which is herein presented in the Appendix.
We close this introduction with a historical remark illustrating the
fascination arising from these densities and their curious geometric features.
H.~Fettis devotes the entire paper \cite{fettis63} to proving that $p_5$ is not
linear on the initial interval $[0,1]$ as ruminated upon by Pearson
\cite{Pea06}. This will be explained in Section~\ref{sec:p5}.
\section{The densities $p_n$}\label{sec:pn}
It is a classical result of Kluyver \cite{Klu06} that $p_n$ has the following
Bessel integral representation:
\begin{equation}\label{eq:pnbessel}
p_n(x) = \int_0^\infty x t J_0(x t)J_0^n(t) \, \md t.
\end{equation}
Here $J_\nu$ is the \emph{Bessel function of the first kind} of order $\nu$.
\begin{remark}
Equation \eqref{eq:pnbessel} naturally generalizes to the case of nonuniform
step lengths. In particular, for $n=2$ and step lengths $a$ and $b$ we
record (see \cite[p. 411]{watson-bessel} or \cite[2.3.2]{hughes-rw}; the
result is attributed to Sonine) that the
corresponding density is
\begin{align}\label{eq:p2ab}
p_2(x;a,b) &= \int_0^\infty x t J_0(x t)J_0(at)J_0(bt) \, \md t \nonumber\\
&= \frac{2x}{\pi \sqrt{( (a+b)^2 - x^2 )( x^2 - (a-b)^2 )}}
\end{align}
for $|a-b| \le x \le a+b$ and $p_2(x;a,b)=0$ otherwise. Observe how \eqref{eq:p2ab}
specializes to \eqref{eq:p2} in the case $a=b=1$.
In the case $n=3$ the density $p_3(x;a,b,c)$ has been evaluated by Nicholson
\cite[p.~414]{watson-bessel} in terms of elliptic integrals directly
generalizing \eqref{eq:p3ell}. The corresponding extensions for four and more
variables appear much less accessible.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
It is visually clear from the graphs in Figure \ref{fig:pn} that $p_n$ is
getting smoother for increasing $n$. This can be made precise from
(\ref{eq:pnbessel}) using the asymptotic formula for $J_0$ for large arguments
and dominated convergence:
\begin{theorem} For each integer $n \ge 0$,
the density $p_{n+4}$ is $\lfloor n/2 \rfloor$ times continuously differentiable.
\end{theorem}
On the other hand, we note from Figure \ref{fig:pn} that the only
points preventing $p_n$ from being smooth appear to be integers.
This will be made precise in Theorem \ref{thm:pnde}.
To this end, we recall a few things about the $s$-th moments $W_n(s)$ of the
density $p_n$ which are given by
\begin{equation}\label{def:Wns}
W_n (s) = \int_0^\infty x^s p_n(x) \, \md x = \int_{[0, 1]^n} \left| \sum_{k = 1}^n e^{2 \pi x_k i}
\right|^s \, \md \tmmathbf{x}.
\end{equation}
Starting with the right-hand side, these moments had been
investigated in \cite{bnsw-rw, bsw-rw2}. There it was shown that
$W_n(s)$ admits an \textit{analytic continuation} to all of the
complex plane with poles of at most order two at certain negative
integers. In particular, $W_3(s)$ has simple poles at $s=-2, -4, -6,
\ldots$ and $W_4(s)$ has double poles at these integers \cite[Thm.
6, Ex. 2 \& 3]{bnsw-rw}.
Moreover, from the combinatorial evaluation
\begin{equation}\label{eq:Wn-even}
W_n (2 k) = \sum_{a_1 + \cdots + a_n = k} \binom{k}{a_1, \ldots, a_n}^2
\end{equation}
for integers $k\ge0$ it followed that $W_n(s)$ satisfies a
functional equation, as in \cite[Ex. 1]{bnsw-rw}, coming from the
inevitable recursion that exists for the right-hand side of
(\ref{eq:Wn-even}) . For instance,
\[ (s + 4)^2 W_3 (s + 4) - 2 (5 s^2 + 30 s + 46) W_3 (s + 2) + 9 (s + 2)^2 W_3 (s) = 0,\]
and equation (\ref{eq:W4fe}) below.
The first part of equation
(\ref{def:Wns}) can be rephrased as saying that $W_n(s-1)$ is the
\emph{Mellin transform} of $p_n$ (\cite{ml-xmellin}). We denote this
by $W_n(s-1)=\MellinS{p_n}{s}$. Conversely, the density $p_n$ is the
\emph{inverse Mellin transform} of $W_n(s-1)$. We intend to exploit
this relation as detailed for $n=4$ in the following example.
\begin{example}[Mellin transforms]\label{eg:p4de}
For $n=4$, the moments $W_4(s)$ satisfy the functional equation
\begin{equation}\label{eq:W4fe}
(s+4)^3 W_4(s+4) - 4(s+3)(5s^2+30s+48) W_4(s+2) + 64(s+2)^3 W_4(s) = 0.
\end{equation}
Recall the following rules for the Mellin transform: if $F (s) = \MellinS{f}{s}$
then in the appropriate strips of convergence
\begin{itemize}
\item $\MellinS{x^{\mu} f (x)}{s} = F (s + \mu)$,
\item $\MellinS{D_x f (x)}{s} = - (s - 1) F (s - 1)$.
\end{itemize}
Here, and below, $D_x$ denotes differentiation with respect to $x$, and, for
the second rule to be true, we have to assume, for instance, that $f$ is
continuously differentiable.
Thus, purely formally, we can translate the functional equation
(\ref{eq:W4fe}) of $W_4$ into the differential equation $A_4 \cdot p_4 (x) =
0$ where $A_4$ is the operator
\begin{align}\label{eq:p4de}
A_4 &= x^4 (\theta+1)^3 - 4x^2 \theta (5\theta^2+3) + 64(\theta-1)^3 \\
&= (x - 4) (x - 2) x^3 (x + 2) (x + 4) D_x^3 + 6 x^4
\left( x^2 - 10 \right) D_x^2 \label{eq:p4deD}\\
&\quad+ x \left( 7 x^4 - 32 x^2 + 64 \right) D_x + \left(
x^2 - 8 \right) \left( x^2 + 8 \right) . \nonumber
\end{align}
Here $\theta=x D_x$.
However, it should be noted that $p_4$ is not continuously differentiable.
Moreover, $p_4(x)$ is approximated by a constant multiple of $\sqrt{4-x}$
as $x\to4^-$ (see Theorem \ref{thm:p4at4}) so that the second derivative of $p_4$ is
not even locally integrable. In particular, it does not have a Mellin
transform in the classical sense.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\begin{theorem}\label{thm:pnde}
Let an integer $n\ge1$ be given.
\begin{itemize}
\item The density $p_n$ satisfies a differential equation of order $n - 1$.
\item If $n$ is even (respectively odd) then $p_n$ is real analytic except
at $0$ and the even (respectively odd) integers $m \leq n$.
\end{itemize}
\end{theorem}
\begin{proof}
As illustrated for $p_4$ in Example \ref{eg:p4de}, we formally use the Mellin
transform method to translate the functional equation of $W_n$ into a
differential equation $A_n \cdot y(x) = 0$. Since $p_n$ is locally integrable
and compactly supported, it has a Mellin transform in the distributional
sense as detailed for instance in \cite{ml-xmellin}. It follows rigorously that $p_n$
solves $A_n \cdot y(x) = 0$ in a distributional sense. In other words, $p_n$
is a weak solution of this differential equation. The degree of this equation
is $n - 1$ because the functional equation satisfied by $W_n$ has
coefficients of degree $n - 1$ as shown in \cite[Thm. 1]{bnsw-rw}.
The leading coefficient of the differential equation (in terms of $D_x$ as in
\eqref{eq:p4deD}) turns out to be
\begin{equation}\label{eq:deleadingcoeff}
x^{n - 1} \prod_{2|(m-n)} (x^2 - m^2)
\end{equation}
where the product is over the even or odd integers $1 \leq m \leq n$
depending on whether $n$ is even or odd. This is discussed below
in Section \ref{sec:deleadingcoeff}.
Thus the leading coefficient of the differential equation is nonzero on $[0,
n]$ except for $0$ and the even or odd integers already mentioned. On each interval
not containing these points it follows, as described for instance in
\cite[Cor. 3.1.6]{hoermander-lpde1}, that $p_n$ is in fact a classical solution of
the differential equation. Moreover the analyticity of the coefficients,
which are polynomials in our case, implies that $p_n$ is piecewise real
analytic as claimed.
\end{proof}
\begin{remark}\label{rk:pnat0}
It is one of the basic properties of the Mellin transform, see for instance
\cite[Appendix B.7]{anacomb}, that the asymptotic behaviour of a function at
zero is determined by the poles of its Mellin transform which lie to the left
of the fundamental strip. It is shown in \cite{bnsw-rw} that the poles of
$W_n(s)$ occur at specific negative integers and are at most of second order.
This translates into the fact that $p_n$ has an expansion at $0$ as a power
series with additional logarithmic terms in the presence of double poles.
This is made explicit in the case of $p_4$ in Example \ref{eg:W4p4mellinat0}.
\end{remark}
\subsection{An explicit recursion}\label{sec:deleadingcoeff}
We close this section by providing details for the claim made in
\eqref{eq:deleadingcoeff}. Recall that the even moments $ f_n(k) := W_n (2 k)
$ satisfy a recurrence of order $\lambda:=\lceil n / 2 \rceil$ with polynomial
coefficients of degree $n - 1$ (see \cite{bnsw-rw}). An entirely explicit
formula for this recurrence is given in \cite{verrill}:
\begin{theorem}\label{thm:verrill}
\begin{equation}\label{eq:recverrill}
\sum_{j\ge0} \left[ k^{n+1} \sum_{\alpha_1,\ldots,\alpha_j}
\prod_{i=1}^j (-\alpha_i)(n+1-\alpha_i) \left( \frac{k-i}{k-i+1} \right)^{\alpha_i-1}
\right] f_n(k-j) = 0
\end{equation}
where the sum is over all sequences $\alpha_1,\ldots,\alpha_j$ such that $0\le\alpha_i\le n$
and $\alpha_{i+1}\le\alpha_i-2$.
\end{theorem}
Observe that \eqref{eq:deleadingcoeff} is easily checked for each fixed $n$ by
applying Theorem \ref{thm:verrill}. We explicitly checked the cases $n\le1000$
(using a recursive formulation of Theorem \ref{thm:verrill} from
\cite{verrill}) while only using this statement for $n\le5$ in this paper. The
fact that \eqref{eq:deleadingcoeff} is true in general is recorded and made
more explicit in Theorem \ref{thm:deleadingcoeffx} below.
For fixed $n$, write the recurrence for $f_n(k)$ in the form
$\sum_{j=0}^{n-1} k^j q_j(K)$ where $q_j$ are polynomials and $K$ is the
shift $k \to k+1$. Then $q_{n-1}$ is the characteristic polynomial of this
recurrence, and, by the rules outlined in Example \ref{eg:p4de}, we find that
the differential equation satisfied by $p_n(x)$ is of the form $ q_{n-1}(x^2)
\theta^{n-1} + \cdots, $ where $\theta = x D_x$ and the dots indicate terms of
lower order in $\theta$.
We claim that the characteristic polynomial of the recurrence
\eqref{eq:recverrill} satisfied by $f_n(k)$ is $ \prod_{2|(m-n)} (x - m^2) $
where the product is over the integers $1 \leq m \leq n$ such that $m\equiv n$
modulo $2$. This implies \eqref{eq:deleadingcoeff}.
By Theorem \ref{thm:verrill} the characteristic polynomial is
\begin{equation}\label{eq:lcverrill}
\sum_{j=0}^\lambda \left[ \sum_{\alpha_1,\ldots,\alpha_j}
\prod_{i=1}^j (-\alpha_i)(n+1-\alpha_i) \right] x^{\lambda-j}
\end{equation}
where $\lambda=\lceil n / 2 \rceil$ and the sum is again over all
sequences $\alpha_1,\ldots,\alpha_j$ such that $0\le\alpha_i\le n$
and $\alpha_{i+1}\le\alpha_i-2$. The claimed evaluation is thus
equivalent to the following identity, first proven by P.~Djakov and
B.~Mityagin \cite{dm1,dm2}. Zagier's more direct and purely
combinatorial proof is given in the Appendix.
\begin{theorem}\label{thm:deleadingcoeffx}
For all integers $n,j\ge1$,
\begin{equation}\label{eq:exconj}
\sum_{ 0 \leqslant m_1, \ldots, m_j < n / 2 \atop m_i < m_{i + 1} }
\prod_{i = 1}^j (n - 2 m_i)^2
= \sum_{ 1 \leqslant \alpha_1, \ldots, \alpha_j \leqslant n \atop
\alpha_i \leqslant \alpha_{i + 1} - 2 }
\prod_{i = 1}^j \alpha_i (n + 1 - \alpha_i).
\end{equation}
\end{theorem}
\section{The density $p_3$}\label{sec:p3}
The elliptic integral evaluation \eqref{eq:p3ell} of $p_3$ is very suitable to
extract information about the features of $p_3$ exposed in Figure \ref{fig:p3}.
It follows, for instance, that $p_3$ has a singularity at $1$. Moreover, using
the known asymptotics for $K(x)$, we may deduce that the singularity is of the
form
\begin{equation}\label{eq:p3at1}
p_3 (x) = \frac{3}{2\pi^2}\log\left( \frac{4}{|x-1|} \right) + O (1)
\end{equation}
as $x\to1$.
We also recall from \cite[Ex. 5]{bsw-rw2} that $p_3$ has the
expansion, valid for $0\le x \le1$,
\begin{equation}\label{eq:p3at0}
p_3(x) = \frac{2 x}{\pi\sqrt{3}}\sum_{k=0}^\infty W_3(2k) \left(\frac x3\right)^{2k}
\end{equation}
where
\begin{equation}
W_3(2k)= \sum_{j=0}^k \binom{k}{j}^2 \binom{2j}{j}
\end{equation}
is the sum of squares of trinomials. Moreover, we have from \cite[Eqn.
29]{bsw-rw2} the functional relation
\begin{equation}\label{eq:p3mod}
p_3(x)=\frac{4x}{(3-x)(x+1)}\,p_3\left(\frac{3-x}{1+x}\right)
\end{equation}
so that \eqref{eq:p3at0} determines $p_3$ completely and also makes apparent the
behaviour at $3$.
We close this section with two more alternative expressions for $p_3$.
\begin{example}[Hypergeometric form for $p_3$]
Using the techniques in \cite{cz-apery} we can deduce from \eqref{eq:p3at0} that
\begin{equation}\label{eq:p3hyp}
p_3(x) =\frac{ 2\,\sqrt{3}x}{\pi \left( 3+{x}^{2} \right)}\,
{_2F_1\left(\frac13,\frac23;\,1;\,{\frac {{x}^{2} \left( 9-x^2 \right)^2}{ \left( 3+x^2 \right)^3}}\right)}
\end{equation}
which is found in a similar but simpler way than the hypergeometric form
of $p_4$ given in Theorem \ref{thm:p4hyp04}.
Once obtained, this identity is easily proven using the differential equation
from Theorem \ref{thm:pnde} satisfied by $p_3$. From \eqref{eq:p3hyp} we
see, for example, that $p_3(\sqrt{3})^2 = \frac{3}{2\pi^2}W_3(-1)$.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\begin{example}[Iterative form for $p_3$]\label{ex:agmp3}
The expression \eqref{eq:p3hyp} can be interpreted in terms of the cubic AGM,
$\tmop{AG}_3$, see \cite{bb-cubicagm}, as follows. Recall that
$\tmop{AG}_3(a,b)$ is the limit of iterating
\begin{equation*}
a_{n+1}=\frac{a_n+2b_n}{3}, \quad b_{n+1}=\sqrt[3]{b_n \left( \frac{a_n^2+a_n b_n+b_n^2}{3} \right)},
\end{equation*}
beginning with $a_0=a$ and $b_0=b$. The iterations converge cubically, thus allowing for very
efficient high-precision evaluation. On the other hand,
\begin{equation*}
\frac{1}{\tmop{AG}_3(1,s)} = \ipFq21{\frac13,\frac23}{1}{1-s^3}
\end{equation*}
so that in consequence of \eqref{eq:p3hyp}, for $0\le x \le3$,
\begin{equation}\label{eq:p3ag3}
p_3(x) = \frac{2\sqrt 3}{\pi} \,\frac{x}{\tmop{AG}_3(3+x^2,3\left|1-x^2\right|^{2/3})}.
\end{equation}
Note that $p_3(3)=\frac{\sqrt{3}}{2\pi}$ is a direct consequence of the final formula.
Finally we remark that the cubic AGM also makes an appearance in the case
$n=4$. We just mention that the modular properties of $p_4$ recorded in
Remark~\ref{rk:p4modularity} can be stated in terms of the theta functions
\begin{equation}
b(\tau) = \frac{\eta(\tau)^3}{\eta(3\tau)}, \quad
c(\tau) = 3\frac{\eta(3\tau)^3}{\eta(\tau)}
\end{equation}
where $\eta$ is the Dedekind eta function defined in \eqref{eq:eta}.
For more information and proper definitions of the functions $b$, $c$ as well
as $a$, which is related by $a^3 = b^3 + c^3$, we refer to \cite{garvan}.
Ultimately we are hopeful that, in search for an analogue of \eqref{eq:p3mod}
for $p_4$, this may lead to an algebraic relation between algebraically
related arguments of $p_4$.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
\section{The density $p_4$}\label{sec:p4}
The densities $p_n$ are recursively related. As in \cite{hughes-rw}, setting
$\phi_{n}(x) = p_{n}(x)/(2\pi x)$, we have that for integers $n\ge2$
\begin{equation}\label{eq:pnrec}
\phi_n(x) = \frac{1}{2\pi}\, \int_0^{2\pi}
\phi_{n-1}\left(\sqrt{x^2-2x \cos \alpha + 1}\right) \, \md \alpha.
\end{equation}
We use this recursive relation to get some quantitative information about the
behaviour of $p_4$ at $x=4$.
\begin{theorem}\label{thm:p4at4}
As $x \rightarrow 4^-$,
\[ p_4(x) = \frac{\sqrt{2}}{\pi^2}\sqrt{4 - x} - \frac{3\sqrt{2}}{16\pi^2} (4-x)^{3/2}
+ \frac{23\sqrt{2}}{512\pi^2} (4-x)^{5/2} + O\left( (4-x)^{7/2} \right). \]
\end{theorem}
\begin{proof}
Set $y=\sqrt{x^2-2x \cos \alpha + 1}$. For $2 < x < 4$,
\[ \phi_4 (x) = \frac{1}{\pi} \int_0^{\pi} \phi_3 (y)\, \mathrm{d} \alpha =
\frac{1}{\pi} \int_0^{\arccos ( \frac{x^2 - 8}{2 x})} \phi_3 (y)\, \mathrm{d} \alpha
\]
since $\phi_3$ is only supported on $[0, 3]$. Note that $\phi_3 (y)$ is
continuous and bounded in the domain of integration. By the Leibniz integral
rule, we can thus differentiate under the integral sign to obtain
\begin{equation}\label{eq:phi4D1}
\phi'_4 (x) = - \frac{1}{\pi} \frac{(x^2 + 8)\, \phi_3 (3)}{x \sqrt{(16 -
x^2) (x^2 - 4)}} + \frac{1}{\pi} \int_0^{\arccos ( \frac{x^2 - 8}{2 x})}
(x - \cos (\alpha)) \,\frac{\phi_3' (y)}{y}\, \mathrm{d} \alpha.
\end{equation}
This shows that $\phi_4'$, and hence $p_4'$, have a singularity at $x =
4$. More specifically,
\[ \phi_4' (x) = - \frac{1}{8 \sqrt{2} \pi^3 \sqrt{4 - x}} + O (1)
\hspace{1em} \text{as $x \rightarrow 4^-$} . \]
Here, we used that $\phi_3 (3) = \frac{\sqrt{3}}{12 \pi^2}$.
It follows that
\[ p_4' (x) = - \frac{1}{\sqrt{2} \pi^2 \sqrt{4 - x}} + O (1) \]
which, upon integration, is the claim to first order. Differentiating
\eqref{eq:phi4D1} twice more proves the claim.
\end{proof}
\begin{remark}
The situation for the singularity at $x = 2^+$ is more complicated since in
(\ref{eq:phi4D1}) both the integral (via the logarithmic singularity of
$\phi_3$ at $1$, see \eqref{eq:p3at1}) and the boundary term contribute.
Numerically, we find, as $x\to2^+$,
\[ p_4' (x) = - \frac{2}{\pi^2 \sqrt{x - 2}} + O (1). \]
On the other hand, the derivative of $p_4$ at 2 from the left is given by
\[ p_4'(2^-) = \frac{\sqrt3}{\pi} \, \pFq32{-\frac12,\frac13,\frac23}{1,1}{1}
- \frac23\, p_4(2). \]
These observations can be proven in hindsight from Theorem \ref{thm:p4hyp}.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
We now turn to the behaviour of $p_4$ at zero which we derive from
the pole structure of $W_4$ as described in Remark \ref{rk:pnat0}.
\begin{figure}[htbp]\label{fig:W45}
\begin{center}
\subfigure[$W_4$\label{fig:W4}]{\includegraphics[width=0.4\textwidth]{W4}}
\hfil
\subfigure[$W_5$\label{fig:W5}]{\includegraphics[width=0.4\textwidth]{W5}}
\caption{$W_4$ and $W_5$ analytically continued to the real line.}
\end{center}
\end{figure}
\begin{example}\label{eg:W4p4mellinat0}
From \cite{bsw-rw2}, we know that $W_4$ has a pole of order $2$ at $- 2$ as
illustrated in Figure \ref{fig:W4}. More specifically, results in Section \ref{sec:Wndiff} give
\[ W_4 (s) = \frac{3}{2 \pi^2} \frac{1}{(s + 2)^2} + \frac{9}{2 \pi^2} \log
(2) \frac{1}{s + 2} + O (1) \]
as $s \rightarrow - 2$. It therefore follows that
\[ p_4 (x) = - \frac{3}{2 \pi^2} x \log (x) + \frac{9}{2 \pi^2} \log (2) x + O(x^3) \]
as $x \rightarrow 0$.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
More generally, $W_4$ has poles of order $2$ at $- 2 k$ for $k$ a positive
integer. Define $s_{4,k}$ and $r_{4,k}$ by
\begin{equation}\label{eq:rs4def}
W_4 (s) = \frac{s_{4, k-1}}{(s + 2 k)^2} + \frac{r_{4, k-1}}{s + 2 k} + O (1)
\end{equation}
as $s \rightarrow - 2 k$. We thus obtain that, as $x \rightarrow 0^+$,
\begin{equation*}
p_4 (x) = \sum_{k = 0}^{K-1} x^{2 k + 1} \left( r_{4, k} - s_{4, k} \log (x)
\right) + O (x^{2 K + 1}).
\end{equation*}
In fact, knowing that $p_4$ solves the linear Fuchsian differential equation
\eqref{eq:p4de} with a regular singularity at $0$ we may conclude:
\begin{theorem}\label{thm:p4at0}
For small values $x>0$,
\begin{equation}\label{eq:p4at0}
p_4 (x) = \sum_{k = 0}^\infty \left( r_{4, k} - s_{4, k} \log (x)
\right)\,x^{2 k + 1}.
\end{equation}
\end{theorem}
Note that
\[ s_{4, k} = \frac{3}{2 \pi^2} \frac{W_4 (2 k)}{8^{2 k}} \]
as the two sequences satisfy the same recurrence and initial conditions. The
numbers $W_4(2k)$ are also known as the Domb numbers (\cite{bbbg-bessel}), and
their generating function in hypergeometric form is given in \cite{rogers-5f4}
and has been further studied in \cite{cz-apery}. We thus have
\begin{equation}\label{eq:p4de-ana}
\sum_{k=0}^\infty s_{4,k} \,x^{2k+1} = \frac{6x}{\pi^2 \left( 4-x^2 \right)}\,
\pFq32{\frac13,\frac12,\frac23}{1,1}{\frac {108x^2}{ \left( x^2-4 \right)^3}}
\end{equation}
which is readily verified to be an analytic solution to the differential
equation satisfied by $p_4$.
\begin{remark}
For future use, we note that \eqref{eq:p4de-ana} can also be written as
\begin{equation}\label{eq:p4de-ana2}
\sum_{k=0}^\infty s_{4,k} \,x^{2k+1} = \frac{24x}{\pi^2 \left( 16-x^2 \right)}\,
\pFq32{\frac13,\frac12,\frac23}{1,1}{\frac {108x^4}{ \left( 16-x^2 \right)^3}}
\end{equation}
which follows from the transformation
\begin{equation}\label{eq:hyptr32}
(1-4x) \pFq32{\frac13,\frac12,\frac23}{1,1}{-\frac{108x}{(1-16x)^3}} =
(1-16x) \pFq32{\frac13,\frac12,\frac23}{1,1}{\frac{108x^2}{(1-4x)^3}}
\end{equation}
given in \cite[(3.1)]{cz-apery}.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
On the other hand, as a consequence of \eqref{eq:rs4def} and the functional
equation \eqref{eq:W4fe} satisfied by $W_4$, the residues $r_{4,k}$ can be
obtained from the recurrence relation
\begin{align}
128 k^3 r_{4,k} &= 4(2k-1)(5k^2-5k+2) r_{4,k-1}
-2(k-1)^3 r_{4,k-2} \nonumber\\
&\quad+3\left( 64k^2 s_{4,k} - (20k^2-20k+6) s_{4,k-1} + (k-1)^2 s_{4,k-2} \right)
\end{align}
with $r_{4,-1}=0$ and $r_{4,0}=\frac{9}{2\pi^2}\log(2)$.
\begin{remark}
In fact, before realizing the connection between the Mellin transform and the
behaviour of $p_4$ at $0$, we empirically found that $p_4$ on $(0,2)$ should
be of the form $r(x)-s(x)\log(x)$ where $a$ and $r$ are odd and analytic. We
then numerically determined the coefficients and observed the relation with
the residues of $W_4$ as given in Theorem \ref{thm:p4at0}.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
The differential equation for $p_4$ has a regular singularity at $0$. A basis
of solutions at $0$ can therefore be obtained via the Frobenius method, see for
instance \cite{ince-ode}. Since the indicial equation has $1$ as a triple root, the
solution \eqref{eq:p4de-ana} is the unique analytic solution at $0$ while the
other solutions have a logarithmic or double logarithmic singularity. The
solution with a logarithmic singularity at $0$ is explicitly given in
\eqref{eq:p4de-logsol}, and, from \eqref{eq:p4at0}, it is clear that $p_4$ on
$(0,2)$ is a linear combination of the analytic and the logarithmic solution.
Moreover, the differential equation for $p_4$ is a symmetric square. In other
words, it can be reduced to a second order differential equation, which after a
quadratic substitution, has 4 regular singularities and is thus of Heun type. In
fact, a hypergeometric representation of $p_4$ with rational argument is possible.
\begin{theorem}\label{thm:p4hyp}
For $2<x<4$,
\begin{equation}\label{eq:p4hyp}
p_4(x) = \frac2{\pi^2}\,\frac{\sqrt {16-x^2}}x\,
\pFq32{\frac12, \frac12, \frac12}{\frac56,\frac76}
{\frac {\left( 16-x^2 \right)^3}{108 x^4}}.
\end{equation}
\end{theorem}
\begin{proof}
Denote the right-hand side of \eqref{eq:p4hyp} by $q_4(x)$ and observe that
the hypergeometric series converges for $2<x<4$. It is routine to verify that
$q_4$ is a solution of the differential equation $A_4 \cdot y(x) = 0$ given
in \eqref{eq:p4de} which is also satisfied by $p_4$ as proven in Theorem
\ref{thm:pnde}. Together with the boundary conditions supplied by Theorem
\ref{thm:p4at4} it follows that $p_4=q_4$.
\end{proof}
We note that Theorem \ref{thm:p4hyp} gives $2\sqrt{16-x^2}/(\pi^2 x)$ as an approximation to $p_4(x)$ near $x=4$, which is much more accurate than the elementary estimates established in Theorem \ref{thm:p4at4}.
\begin{corollary}
In particular,
\begin{equation}
p_4(2)=\frac {2^{7/3}\pi}{3\sqrt{3}}\,\Gamma\left (\frac 23 \right)^{-6} = \frac{\sqrt 3}{\pi}\,W_3(-1).
\end{equation}
\end{corollary}
Quite marvelously, as first discovered numerically:
\begin{theorem}\label{thm:p4hyp04}
For $0<x<4$,
\begin{equation}\label{eq:p4hyp04}
p_4(x) = \frac2{\pi^2}\,\frac{\sqrt {16-x^2}}x\,
\Re \pFq32{\frac12, \frac12, \frac12}{\frac56,\frac76}
{\frac {\left( 16-x^2 \right)^3}{108 x^4}}.
\end{equation}
\end{theorem}
\begin{proof}
To obtain the analytic continuation of the ${}_3F_2$ for $0<x<2$ we employ
the formula \cite[5.3]{luke-sf}, valid for all $z$,
\begin{align*}
\pFq{q+1}{q}{a_1,\ldots,a_{q+1}}{b_1,\ldots,b_q}{z} &= \frac{\prod_j\Gamma(b_j)}{\prod_j\Gamma(a_j)}
\sum_{k=1}^{q+1} \frac{\Gamma(a_k)\prod_{j\ne k}\Gamma(a_j-a_k)}{\prod_j\Gamma(b_j-a_k)} (-z)^{-a_k} \\
&\quad \times\pFq{q+1}{q}{a_k,\{a_k-b_j+1\}_j}{\{a_k-a_j+1\}_{j\ne k}}{\frac1z},
\end{align*}
which requires the $a_j$ to not differ by integers. Therefore we apply it to
\begin{equation*}
\pFq32{\frac12+\varepsilon,\frac12,\frac12-\varepsilon}{\frac56,\frac76}{z}.
\end{equation*}
and take the limit as $\varepsilon\to0$. This ultimately produces, for $z>1$,
\begin{align}\label{eq:p4de-logsol}
\Re \pFq32{\frac12, \frac12, \frac12}{\frac56, \frac76}{z}
&= \frac{\log(108z)}{2\sqrt{3z}}\, \pFq32{\frac13, \frac12, \frac23}{1, 1}{\frac{1}{z}}\\
&\quad+ \frac{1}{2\sqrt{3z}}\,\sum_{n=0}^\infty \frac{(\tfrac13)_n (\tfrac12)_n
(\tfrac23)_n}{ n!^3} \left( \frac1z \right)^n (5H_n-2H_{2n}-3H_{3n}).\nonumber
\end{align}
Here $H_n = \sum_{k=1}^n 1/k$ is the $n$-th harmonic number. Now, insert the
appropriate argument for $z$ and the factors so the left-hand side
corresponds to the claimed closed form. Observing that
\begin{equation*}
\left(\tfrac13\right)_n \left(\tfrac12\right)_n \left(\tfrac23\right)_n = \frac{(2n)!(3n)!}{108^n (n!)^2},
\end{equation*}
we thus find that the right-hand side of \eqref{eq:p4hyp04} is given by
$-\log(x) S_4(x)$ plus
\begin{align*}
\frac{6}{\pi^2} \sum_{n=0}^\infty \frac{(2n)!(3n)!}{(n!)^5}
\frac{x^{4n+1}}{(16-x^2)^{3n}} \left( 5H_n-2H_{2n}-3H_{3n} +3\log(16-x^2) \right) \nonumber
\end{align*}
where $S_4$ is the solution (analytic at $0$) to the differential equation
for $p_4$ given in \eqref{eq:p4de-ana2}. This combination can now be verified
to be a formal and hence actual solution of the differential equation for
$p_4$. Together with the boundary conditions supplied by Theorem
\ref{thm:p4at0} this proves the claim.
\end{proof}
\begin{remark}\label{rk:p4hypform}
Let us indicate how the hypergeometric expression for $p_4$ given in
Theorem~\ref{thm:p4hyp} was discovered. Consider the generating series
\begin{equation}\label{eq:y0def}
y_0(z) = \sum_{k=0}^\infty W_4(2k) z^k
\end{equation}
of the Domb numbers which is just a rescaled version of \eqref{eq:p4de-ana}.
Corresponding to \eqref{eq:p4de-ana2}, the hypergeometric form for this
series given in \cite{rogers-5f4} is
\begin{equation}\label{eq:y0hyp}
y_0(z) = \frac{1}{1-4z}\,\pFq32{\frac13,\frac12,\frac23}{1,1}{\frac{108z^2}{(1-4z)^3}}
\end{equation}
which converges for $|z|<1/16$.
$y_0$ satisfies the differential equation $B_4 \cdot y_0(z) = 0$ where
\begin{equation}\label{eq:B4}
B_4 = 64z^2 (\theta+1)^3 - 2z (2\theta+1)(5\theta^2+5\theta+2) + \theta^3
\end{equation}
and $\theta = z \frac{\mathrm{d}}{\mathrm{d} z}$. Up to a change of variables this is
\eqref{eq:p4de}; $y_0$ is the unique solution which is analytic at zero and
takes the value $1$ at zero; the other solutions which are not a multiple of
$y_0$ have a single or double logarithmic singularity. Let $y_1$ be the
solution characterized by
\begin{equation}\label{eq:y1def}
y_1(z) - y_0(z) \log(z) \in z \mathbb{Q}[[z]].
\end{equation}
Note that it follows from \eqref{eq:y1def} as well as Theorem~\ref{thm:p4at0}
together with the initial values $s_{4,0} = \frac3{2\pi^2}$ and $r_{4,0} =
s_{4,0} \log(8)$ that $p_4$, for small positive argument, is given by
\begin{equation}\label{eq:p4y1}
p_4(x) = - \frac{3x}{4\pi^2} \; y_1\left( \frac{x^2}{64} \right).
\end{equation}
If $x\in(2,4)$ and $z=x^2/64$ then the argument $t=\frac{108z^2}{(1-4z)^3}$
of the hypergeometric function in \eqref{eq:y0hyp} takes the values
$(1,\infty)$. We therefore consider the solutions of the corresponding
hypergeometric equation at infinity. A standard basis for these is
\begin{align}
t^{-1/3}\pFq32{\frac13,\frac13,\frac13}{\frac23,\frac56}{\frac1t}, \quad
t^{-1/2}\pFq32{\frac12,\frac12,\frac12}{\frac56,\frac76}{\frac1t}, \quad
t^{-2/3}\pFq32{\frac23,\frac23,\frac23}{\frac43,\frac76}{\frac1t}.
\end{align}
In fact, the second element suffices to express $p_4$ on the interval $(2,4)$
as shown in Theorem~\ref{thm:p4hyp}.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
We close this section by showing that, remarkably, $p_4$ has modular
structure.
\begin{remark}\label{rk:p4modularity}
As shown in \cite{cz-apery} the series $y_0$ defined in
\eqref{eq:y0def} possesses the modular parameterization
\begin{equation}\label{eq:y0modular}
y_0\left( -\frac{\eta(2\tau)^6\eta(6\tau)^6}{\eta(\tau)^6\eta(3\tau)^6} \right)
= \frac{\eta(\tau)^4\eta(3\tau)^4}{\eta(2\tau)^2\eta(6\tau)^2}.
\end{equation}
Here $\eta$ is the \emph{Dedekind eta function} defined as
\begin{equation}\label{eq:eta}
\eta(\tau) = q^{1/24}\, \prod_{n=1}^\infty(1-q^n)
= q^{1/24}\,\sum_{n=-\infty}^\infty (-1)^n q^{n(3n+1)/2},
\end{equation}
where $q=e^{2\pi i\tau}$.
Moreover, the quotient of the logarithmic solution $y_1$ defined in
\eqref{eq:y1def} and $y_0$ is related to the modular parameter $\tau$ used in
\eqref{eq:y0modular} by
\begin{equation}\label{eq:y1y0}
\exp\left( \frac{y_1(z)}{y_0(z)} \right) =e^{(2\tau+1)\pi i} = -q.
\end{equation}
Combining \eqref{eq:y0modular}, \eqref{eq:y1y0} and \eqref{eq:p4y1} one
obtains the modular representation
\begin{equation}\label{eq:p4modular}
p_4\left( 8i\frac{\eta(2\tau)^3\eta(6\tau)^3}{\eta(\tau)^3\eta(3\tau)^3} \right)
= \frac{6(2\tau+1)}{\pi} \eta(\tau)\eta(2\tau)\eta(3\tau)\eta(6\tau)
\end{equation}
valid when the argument of $p_4$ is small and positive. This is the case for
$\tau = -1/2 + i y$ when $y>0$. Remarkably, the argument attains the value
$1$ at the quadratic irrationality $\tau = (\sqrt{-5/3}-1)/2$ (the $5/3$rd
singular value of the next section). As a consequence, the value $p_4(1)$
has a nice evaluation which is given in Theorem~\ref{thm:r50cs}.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
\section{The density $p_5$}\label{sec:p5}
As shown in \cite{bsw-rw2}, $W_5 (s)$ has simple poles at $- 2, - 4,
\ldots$, compare Figure \ref{fig:W5}. We write $r_{5, k} = \tmop{Res}_{- 2
k - 2} W_5$ for the residue of $W_5$ at $s=-2k-2$. A surprising
bonus is an evaluation of $r_{5,0}=p_4(1) \approx 0.3299338011$, the
residue at $s=-2$. This is because in general for $n \ge 4$, one has
$$\tmop{Res}_{- 2} W_{n+1}=p_{n+1}'(0)=p_n(1),$$ as follows from
\cite[Prop. 1(b)]{bsw-rw2}; here $p_{n+1}'(0)$ denotes the
derivative from the right at zero.
Explicitly, using Theorem \ref{thm:p4hyp04}, we have,
\begin{equation}\label{eq:res50}
r_{5,0} =p_5'(0) = \frac{2\sqrt{15}}{\pi^2}\,
{\rm Re}\,{\pFq32{\frac12, \frac12, \frac12}{\frac56,\frac76}{\frac{125}4}}.
\end{equation}
In fact, based on the modularity of $p_4$ discussed in
Remark~\ref{rk:p4modularity} we find:
\begin{theorem}\label{thm:r50cs}
\begin{equation}\label{eq:r50cs}
r_{5,0} =
\frac{1}{2\pi^2}
\sqrt{\frac{\Gamma(\frac{1}{15})\Gamma(\frac{2}{15})\Gamma(\frac{4}{15})\Gamma(\frac{8}{15})}
{5\Gamma(\frac{7}{15})\Gamma(\frac{11}{15})\Gamma(\frac{13}{15})\Gamma(\frac{14}{15})}}.
\end{equation}
\end{theorem}
\begin{proof}
The value $\tau = (\sqrt{-5/3}-1)/2$ in \eqref{eq:p4modular} gives the value
$p_4(1) = r_{5,0}$. Applying the Chowla--Selberg formula
\cite{sc67,borwein-piagm} to evaluate the eta functions yields the claimed
evaluation.
\end{proof}
Using \cite[Table 4, (ii)]{bz92},
\eqref{eq:r50cs} may be simplified to
\begin{align}\label{eq:r50g2}
r_{5,0} &= \frac{\sqrt {5}}{40}\,\frac{\Gamma(\frac{1}{15})\Gamma(\frac{2}{15})\Gamma(\frac{4}{15})\Gamma(\frac{8}{15})}{\pi^4 } \\
&= \frac{3\sqrt {5} }{\pi^3}\,\frac{\left( \sqrt {5}-1 \right)}{2} \,K^2_{15} = \frac{\sqrt{15}}{\pi^3} \,K_{5/3} K_{15},
\end{align}
where $K_{15}$ and $K_{5/3}$ are the complete elliptic integral at the 15th and
$5/3$rd singular values \cite{borwein-piagm}.
Remarkably, these evaluations appear to extend to
$r_{5,1}\approx0.006616730259$, the residue at $s=-4$. Resemblance
to \emph{the tiny nome of Bologna} \cite{bbbg-bessel} led us to
discover
--- and then check to 400 places using \eqref{r50} and \eqref{r51} --- that
\begin{align}\label{eq:r51g}
r_{5,1} &\stackrel{?}{=} {\frac {13}{1800\sqrt {5}}}\,{\frac {\Gamma(\frac{1}{15})\Gamma(\frac{2}{15})\Gamma(\frac{4}{15})\Gamma(\frac{8}{15})}{{\pi }^{4}}}
-\frac1{\sqrt {5}}\,{\frac {\Gamma(\frac{7}{15})\Gamma(\frac{11}{15})\Gamma(\frac{13}{15})\Gamma(\frac{14}{15})}{{\pi }^{4}}}.
\end{align}
Using \eqref{eq:r50g2} this evaluation can be neatly restated as
\begin{align}
r_{5,1} \stackrel{?}{=} \frac{13}{225} r_{5,0} -\frac2{5\pi^4}\frac{1}{r_{5,0}}.
\end{align}
We summarize our knowledge as follows:
\begin{theorem}\label{thm:p5}
The density $p_5$ is real analytic on $(0,5)$ except at $1$ and $3$
and satisfies the differential equation $A_5 \cdot p_5(x) = 0$
where $A_5$ is the operator
\begin{align}\label{eq:p5de}
A_5 &= x^6 (\theta+1)^4
- x^4 (35\theta^4 + 42\theta^2 + 3) \\
&\quad+ x^2 (259(\theta-1)^4 + 104(\theta-1)^2)
- \left( 15(\theta-3)(\theta-1) \right)^2 \nonumber
\end{align}
and $\theta=x D_x$.
Moreover, for small $x>0$,
\begin{equation}\label{eq:p5at0}
p_5 (x) = \sum_{k = 0}^\infty r_{5, k}\, x^{2 k + 1}
\end{equation}
where
\begin{align}\label{eq:res5rec}
\left( 15 (2k+2) (2k+4) \right)^2 r_{5,k+2}
&= \left( 259(2k+2)^4 + 104(2k+2)^2 \right) r_{5,k+1} \nonumber\\
&\quad- \left( 35(2k+1)^4 + 42(2k+1)^2 + 3 \right) r_{5,k}
+ (2k)^4 r_{5,k-1}
\end{align}
with explicit initial values $r_{5,-1}=0$ and $r_{5,0}$, $r_{5,1}$
given by \eqref{eq:r50g2} and \eqref{eq:r51g} above.
\end{theorem}
\begin{proof}
First, the differential equation \eqref{eq:p5de} is computed as was
that for $p_4$, see \eqref{eq:p4de}. Next, as detailed in \cite[Ex.
3]{bsw-rw2} the residues satisfy the recurrence relation
\eqref{eq:res5rec} with the given initial values. Finally,
proceeding as for \eqref{eq:p4at0}, we deduce that \eqref{eq:p5at0}
holds for small $x>0$.
\end{proof}
Numerically, the series \eqref{eq:p5at0} appears to converge for $|x|<3$ which
is in accordance with $\tfrac19$ being a root of the characteristic
polynomial of the recurrence \eqref{eq:res5rec}; see also
\eqref{eq:deleadingcoeff}. The series \eqref{eq:p5at0} is depicted
in Figure \ref{fig:p5x}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=0.5\textwidth]{p5x}
\caption{The series \eqref{eq:p5at0} (dotted) and $p_5$.}
\label{fig:p5x}
\end{center}
\end{figure}
Since the poles of $W_5$ are simple, no logarithmic terms are
involved in \eqref{eq:p5at0} as opposed to \eqref{eq:p4at0}. In particular, by
computing a few more residues from \eqref{eq:res5rec},
\[ p_5(x) = 0.329934\,x + 0.00661673\,{x}^{3} + 0.000262333\,{x}^{5} + 0.0000141185\,{x}^{7} + O(x^9) \]
near 0 (with each coefficient given to six digits of precision only),
explaining the strikingly straight shape of $p_5(x)$ on $[0,1]$. This
phenomenon was observed by Pearson \cite{Pea06} who stated that for $p_5(x)/x$
between $x=0$ and $x=1$,
\begin{quote}
``the graphical construction, however carefully reinvestigated, did not
permit of our considering the curve to be anything but a \emph{straight}
line\ldots Even if it is not absolutely true, it exemplifies the
extraordinary power of such integrals of $J$ products [that is,
\eqref{eq:pnbessel}] to give extremely close approximations to such simple
forms as horizontal lines.''
\end{quote}
This conjecture was investigated in detail in \cite{fettis63} wherein the
nonlinearity was first rigorously established. This work and various more
recent papers highlight the difficulty of computing the underlying Bessel
integrals.
\begin{remark}
Recall from Example \ref{eg:W4p4mellinat0} that the asymptotic behaviour of
$p_n$ at zero is determined by the poles of the moments $W_n(s)$. To obtain
information about the behaviour of $p_n(x)$ as $x \to n^-$, we consider the
``reversed'' densities $\tilde{p}_n(x) = p_n(n-x)$ and their moments
$\tilde{W}_n(s)$. For non-negative integers $k$,
\begin{equation*}
\tilde{W}_n(k) = \int_0^n x^k \tilde{p}_n(x) \, \md x
= \int_0^n (n-x)^k p_n(x) \, \md x
= \sum_{j=0}^k \binom{k}{j} (-1)^j n^{k-j} W_n(j)
\end{equation*}
On the other hand, we can find a recurrence satisfied by the $\tilde{W}_n(s)$
as follows: a differential equation for the densities $\tilde{p}_n(x)$ is
obtained from Theorem \ref{thm:pnde} by a change of variables. The Mellin
transform method as described in Example \ref{eg:p4de} then provides a
recurrence for the moments $\tilde{W}_n(s)$. We next apply the same
reasoning as in \cite{bsw-rw2} to obtain information about the pole structure of
$\tilde{W}_n(s)$. It should be emphasized that this involves knowledge about
initial conditions in term of explicit values of initial moments $W_n(2k)$.
For instance, in the case $n=4$, we find that the moments $\tilde{W}_4(s)$
have simple poles at $-\tfrac32, -\tfrac52, -\tfrac72, \ldots$ which
predicts an expansion of $p_4(x)$ as given in Theorem \ref{thm:p4at4}.
For $n=5$, we learn that $\tilde{W}_5(s)$ has simple poles at
$s=-2,-3,-4,\ldots$. It then follows, as for \eqref{eq:p5at0}, that $p_5 (x) =
\sum_{k = 0}^\infty \tilde{r}_{5, k}\, (x-5)^{k + 1}$ for $x\le5$ and close
to $5$. The $\tilde{r}_{5,k}$ are the residues of $\tilde{W}_5(s)$ at
$s=-k-2$.
\hspace*{\fill}$\Diamond$\medskip
\end{remark}
\section{Derivative evaluations of $W_n$}\label{sec:Wndiff}
As illustrated by Theorem \ref{thm:p4at0}, the residues of $W_n(s)$ are very
important for studying the densities $p_n$ as they directly translate into
behaviour of $p_n$ at $0$. The residues may be obtained as a linear combination of
the values of $W_n(s)$ and $W_n'(s)$.
\begin{example}[Residues of $W_n$]
Using the functional equation for $W_3(s)$ and L'H\^{o}pital's rule we find that
the residue at $s=-2$ can be expressed as
\begin{equation}\label{eq:W3res}
\tmop{Res}_{-2}(W_3) = \frac{8+12W_3'(0)-4W_3'(2)}9.
\end{equation}
This is a general principle and we likewise obtain for instance:
\begin{align}\label{r50}
\tmop{Res}_{-2}(W_5) &= \frac{16+1140W_5'(0)-804W_5'(2)+64W_5'(4)}{225},\\\label{r51}
\tmop{Res}_{-4}(W_5) &= \frac{26 \tmop{Res}_{-2}(W_5) -16 -20 W_5'(0) + 4 W_5'(2)}{225}.
\end{align}
In the presence of double poles, as for $W_4$,
\begin{equation}\label{w4coeff}
\lim_{s \to -2}(s+2)^2 W_4(s) = \frac{3+4W_4'(0)-W_4'(2)}8
\end{equation}
and for the residue:
\begin{equation}\label{w4resid}
\tmop{Res}_{-2}(W_4) = \frac{9+18W_4'(0)-3W_4'(2)+4W_4''(0)-W_4''(2)}{16}.
\end{equation}
Equations (\ref{w4coeff}, \ref{w4resid}) are used in Example \ref{eg:W4p4mellinat0} and
each unknown is evaluated below.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
We are therefore interested in evaluations of the derivatives of $W_n$ for even
arguments.
\begin{example}[Derivatives of $W_3$ and $W_4$]
Differentiating the double integral for $W_3(s)$ (\ref{def:Wns}) under
the integral sign, we have
\begin{equation*}
W_3'(0) = \frac12 \int_0^1 \int_0^1 \log(4\sin(\pi y)\cos(2\pi x)+3-2\cos(2\pi y)) \, \md x \, \md y.
\end{equation*}
Then, using \[\int_0^1 \log(a+b\cos(2\pi x)) \, \md x =
\log\left(\frac12\left(a+\sqrt{a^2-b^2}\right)\right) \ \mathrm{for} \ a >b >0,\] we deduce
\begin{equation}\label{w3-cl}
W_3'(0) = \int_{1/6}^{5/6} \log(2\sin(\pi y)) \, \md y = \frac{1}{\pi}\,\tmop{Cl}\left(\frac{\pi}{3}\right),
\end{equation}
where $\tmop{Cl}$ denotes the \emph{Clausen} function. Knowing as we do that the
residue at $s=-2$ is $2/(\sqrt{3}\pi)$, we can thus also obtain from (\ref{eq:W3res}) that
\[ W_3'(2)=2+ \frac{3}{\pi} \tmop{Cl}\left(\frac{\pi}{3}\right) - \frac{3\sqrt{3}}{2\pi}. \]
In like fashion,
\begin{align}\label{w4diffm1}
W_4'(0) & = \frac 3{8\pi^2} \int_0^\pi \int_0^\pi
\log\left(3+2\,\cos x+2\,\cos y +2\,\cos(x-y) \right) \, \md x \, \md y \nonumber\\
& = \frac{7}{2}\frac{\zeta(3)}{\pi^2}.
\end{align}
The final equality will be shown in Example \ref{ex:w340}. Note that we may
also write
\begin{equation*}
W_3'(0) = \frac{1}{8\pi^2} \int_0^{2\pi}\int_0^{2\pi}
\log(3+2\cos x+2\cos y+2\cos(x-y)) \, \md x \, \md y.
\end{equation*}
The similarity between $W_3'(0)$ and $W_4'(0)$ is not coincidental, but comes
from applying
\begin{equation*}
\int_0^1 \log\left((a+\cos 2\pi x)^2 + (b+\sin 2\pi x)^2\right) \, \md x
= \left\{ \begin{array}{ll}
\log(a^2+b^2) & \text{if $a^2+b^2>1$,} \\
0 & \text{otherwise}
\end{array} \right.
\end{equation*}
to the triple integral of $W_4'(0)$. As this reduction breaks the symmetry,
we cannot apply it to $W_5'(0)$ to get a similar integral.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
In general, differentiating the Bessel integral expression
\begin{equation}\label{eq-broadhurst}
W_n(s) = 2^{s+1-k} \frac{\Gamma(1+\tfrac{s}{2})}{\Gamma(k-\tfrac{s}{2})}
\int_0^\infty x^{2k-s-1} \left(-\frac{1}{x}\frac{\mathrm{d}}{\mathrm{d} x}\right)^k J_0^n(x) \, \md x,
\end{equation}
obtained by David Broadhurst \cite{broadhurst-rw} and discussed in \cite{bsw-rw2},
under the integral sign gives
\begin{align}\label{eq:wnd0}
W_n'(0) & = n \int_0^\infty \left(\log\left(\frac 2x \right)-\gamma\right)
J_0^{n-1}(x)J_1(x) \, \md x \nonumber\\
& = \log(2)-\gamma-n \int_0^\infty \log(x)J_0^{n-1}(x)J_1(x) \, \md x,
\end{align}
where $\gamma$ is the \emph{Euler-Mascheroni} constant, and
\begin{equation*}
W_n''(0) = n \int_0^\infty \left(\log\left(\frac 2x \right)-\gamma\right)^2
J_0^{n-1}(x)J_1(x) \, \md x.
\end{equation*}
Likewise
\begin{equation*}
W_n'(-1) = (\log(2) -\gamma)W_n(-1)-\int_0^\infty \log(x) J_0^n(x) \, \md x,
\end{equation*}
and
\begin{equation*}
W_n'(1) = \int_0^\infty \frac{n}{x} J_0^{n-1}(x)J_1(x)\left(1-\gamma-\log(2x)\right) \, \md x.
\end{equation*}
\begin{remark}
We may therefore obtain many identities by comparing the above equations to
known values. For instance,
\begin{equation*}
3 \int_0^\infty \log(x)J_0^2(x)J_1(x) \, \md x
= \log(2)-\gamma-\frac{1}{\pi}\tmop{Cl}\left( \frac{\pi}{3}\right).
\end{equation*}
\vskip-\baselineskip\hspace*{\fill}$\Diamond$\medskip
\end{remark}
\begin{example}[Derivatives of $W_5$]
In the case $n=5$,
\begin{equation*}
W_5'(0) = 5 \int_0^\infty \left( \log \left( \frac 2t \right) -\gamma \right)
J_0^4(t) J_1(t) \, \md t \approx 0.54441256
\end{equation*}
with similar but more elaborate formulae for $W_5'(2)$ and $W_5'(4)$.
Observe that in general we also have
\begin{equation}
W_n'(0) = \log(2)-\gamma-\int_0^1 \left(J_0^n(x)-1\right) \frac{\mathrm{d} x}{x}
- \int_1^\infty J_0^n(x) \frac{\mathrm{d} x}{x},
\end{equation}
which is useful numerically. \hspace*{\fill}$\Diamond$\medskip
\end{example}
In fact, the hypergeometric representation of $W_3$ and $W_4$ first obtained
in \cite{crandall-rw} and recalled below also makes derivation of the derivatives
of $W_3$ and $W_4$ possible.
\begin{corollary}[Hypergeometric forms]\label{cor:W34hyp}
For $s$ not an odd integer, we have
\begin{align}
W_3(s) &= \frac{1}{2^{2s+1}} \tan\left(\frac{\pi s}{2}\right)
\binom{s}{\frac{s-1}{2}}^2
\pFq32{\frac12, \frac12, \frac12}{\frac{s+3}{2}, \frac{s+3}{2}}{\frac14}
+ \binom{s}{\frac{s}{2}} \pFq32{-\frac s2, -\frac s2, -\frac s2}{1, -\frac{s-1}{2}}{\frac14}, \label{eq:cf3s}\\
\shortintertext{and, if also $\Re s > -2$, we have}
W_4(s) &= \frac{1}{2^{2s}} \tan\left(\frac{\pi s}{2}\right)
\binom{s}{\frac{s-1}{2}}^3
\pFq43{\frac12, \frac12, \frac12 , \frac s2 +1}{\frac{s+3}{2},\frac{s+3}{2},\frac{s+3}{2}}{1}
+ \binom{s}{\frac{s}{2}} \pFq43{\frac12,-\frac s2, -\frac s2, -\frac s2}{1,1,-\frac{s-1}{2}}{1}. \label{eq:cf4s}
\end{align}
\end{corollary}
\begin{example}[Evaluation of $W_3'(0)$ and $W_4'(0)$] \label{ex:w340}
If we write (\ref{eq:cf3s}) or (\ref{eq:cf4s}) as
$W_n(s)=f_1(s)F_1(s)+f_2(s)F_2(s)$, where $F_1, F_2$ are the corresponding
hypergeometric functions, then it can be readily verified that $f_1(0) =
f_2'(0) = F_2'(0) =0$. Thus, differentiating (\ref{eq:cf3s}) by appealing to
the product rule we get:
\begin{equation*}
W_3'(0) = \frac{1}{\pi} \pFq32{\frac12, \frac12, \frac12}{\frac32, \frac32}{\frac14}
= \frac{1}{\pi} \tmop{Cl} \left( \frac{\pi}{3} \right).
\end{equation*}
The last equality follows from setting $\theta=\pi/6$ in the identity
\begin{equation*}
2\sin(\theta) \pFq32{\frac12, \frac12, \frac12}{\frac32, \frac32}{\sin^2 \theta}
= \tmop{Cl}\left( 2\,\theta \right) +2\,\theta \log \left( 2\sin \theta \right).
\end{equation*}
Likewise, differentiating (\ref{eq:cf4s}) gives
\begin{equation}
W_4'(0) = \frac{4}{\pi^2} \pFq43{\frac12, \frac12, \frac12, 1}{\frac32, \frac32, \frac32}{1}
= \frac{7 \zeta(3)}{2\pi^2},
\end{equation}
thus verifying (\ref{w4diffm1}). In this case the hypergeometric
evaluation
\begin{equation*}
\pFq43{\frac12, \frac12, \frac12, 1}{\frac32, \frac32, \frac32}{1}
= \sum_{n=0}^\infty \frac{1}{(2n+1)^3}= \frac 78 \zeta(3),
\end{equation*}
is elementary.
\hspace*{\fill}$\Diamond$\medskip
\end{example}
Differentiating (\ref{eq:cf3s}) at $s=2$ leads to the evaluation
\begin{equation*}
\pFq32{\frac12, \frac12, \frac12}{\frac52, \frac52}{\frac14}
= \frac{27}{4}\left(\tmop{Cl}\left(\frac\pi 3\right)-\frac{\sqrt{3}} 2\right),
\end{equation*}
while from (\ref{eq:cf4s}) at $s=2$ we obtain
\begin{equation}
W_4'(2) = 3 + \frac{14 \zeta(3) - 12}{\pi^2}.
\end{equation}
Thus we have enough information to evaluate (\ref{w4coeff}) (with the answer $3/(2\pi^2)$).
Note that with two such starting values, all derivatives of $W_3(s)$ or $W_4(s)$
at even $s$ may be computed recursively.
We also note here that the same technique yields
\begin{align}\label{eq:w3dd}
W_3''(0) & = \frac{\pi^2}{12} - \frac{2}{\pi} \sum_{n=0}^\infty
\frac{\binom{2n}{n}}{4^{2n}}\frac{ H_{n+1/2}}{(2n+1)^2} \\
& = \frac{\pi^2}{12} + \frac{4 \log (2)}{\pi}\tmop{Cl}\left(\frac\pi 3\right)
- \frac{4}{\pi} \sum_{n=0}^\infty \frac{\binom{2n}{n}}{4^{2n}}
\frac{\sum_{k=0}^n \frac 1{2k+1}}{(2n+1)^2},
\end{align}
and, quite remarkably,
\begin{align}\label{eq:w4dd}
W_4''(0) & = \frac{\pi^2}{12}+\frac{7\zeta(3)\log (2)}{\pi^2}
+ \frac{4}{\pi^2} \sum_{n=0}^\infty \frac{H_n-3H_{n+1/2}}{(2n+1)^3} \\
& = \frac{24{\rm Li}_4\left(\frac 12\right)-18\zeta(4)+21\zeta(3) \log (2)
- 6\zeta(2)\log^2 (2) +\log^4 (2)}{\pi^2},\nonumber
\end{align}
where the very final evaluation is obtained from results in \cite[\S5]{bzb}.
Here ${\rm Li}_4(1/2)$ is the \emph{polylogarithm} of order 4, while $H_n
:=\gamma+\Psi(n+1)$ denotes the $n$th harmonic number, where $\Psi$ is the
\emph{digamma} function. So for non-negative integers $n$, we
have explicitly $H_n= \sum_{k=1}^n 1/k$, as before, and
\begin{equation*}
H_{n+1/2}= 2\sum_{k=1}^{n+1} \frac1{2k-1}-2\log (2).
\end{equation*}
An evaluation of $W_3''(0)$ in terms of polylogarithmic constants is given in
\cite{logsin1}. In particular, this gives an evaluation of the sum on the
right-hand side of \eqref{eq:w3dd}.
Finally, the corresponding sum for $W_4''(2)$ may be split into a telescoping
part and a part involving $W_4''(0)$. Therefore, it can also be written as a
linear combination of the constants used in (\ref{eq:w4dd}). In summary, we have
all the pieces to evaluate (\ref{w4resid}), obtaining the answer $9\log
(2)/(2\pi^2)$.
\subsection{Relations with Mahler measure}
For a (Laurent) polynomial $f (x_1, \ldots, x_n)$, its \emph{logarithmic Mahler
measure}, see for instance \cite{rtv-mahler}, is defined as
\[ m (f) = \int_0^1 \ldots \int_0^1 \log \left| f \left( e^{2 \pi i t_1},
\ldots, e^{2 \pi i t_n} \right) \right| \mathrm{d} t_1 \cdots \mathrm{d} t_n. \]
Recall that the $s$th moments of an $n$-step random walk are given by
\[ W_n (s) = \int_0^1 \ldots \int_0^1 \left| \sum_{k = 1}^n e^{2 \pi
i t_k} \right|^s \mathrm{d} t_1 \cdots \mathrm{d} t_n =\|x_1 +
\ldots + x_n \|_s^s \]
where $\| \cdot \|_p$ denotes the $p$-norm over the unit $n$-torus, and hence
\[ W_n' (0) = m (x_1 + \ldots + x_n) = m (1 + x_1 + \ldots + x_{n-1}) . \]
Thus the derivative evaluations in the previous sections are also Mahler
measure evaluations. In particular, we rediscovered
\[ W_3' (0) = \frac{1}{\pi} \tmop{Cl} \left( \frac{\pi}{3} \right) = L'
(\chi_{- 3}, - 1) = m (1 + x_1 + x_2), \]
along with
\[ W_4' (0) = \frac{7 \zeta (3)}{2 \pi^2} = m (1 + x_1 + x_2 + x_3) \]
which are both due to C. Smyth \cite[(1.1) and (1.2)]{rtv-mahler} with proofs first published in \cite[Appendix 1]{boyd}.
With this connection realized, we find the following conjectural expressions
put forth by Rodriguez-Villegas, mentioned in different form in \cite{finch},
\begin{equation}\label{eq:vil1} W_5'(0) \stackrel{?}{=} \left(\frac{15}{4\pi^2}\right)^{5/2}\,
\int_0^\infty \left\{\eta^3(e^{-3t})\eta^3(e^{-5t})
+\eta^3(e^{-t})\eta^3(e^{-15t})\right\} t^3\,\mathrm{d} t \end{equation}
and
\begin{equation}\label{eq:vil2} W_6'(0) \stackrel{?}{=} \left(\frac{3}{\pi^2}\right)^{3}\,
\int_0^\infty \,\eta^2(e^{-t})\eta^2(e^{-2t}) \eta^2(e^{-3t})\eta^2(e^{-6t})\,t^4\,\mathrm{d} t, \end{equation}
where $\eta$ was defined in~\eqref{eq:eta}. We have confirmed numerically that
the evaluation of $W_5'(0)$ in (\ref{eq:vil1}) holds to $600$ places.
Likewise, we have confirmed that (\ref{eq:vil2}) holds to $80$ places. Details
of these somewhat arduous confirmations are given in \cite{bb-combat}.
Differentiating the series expansion for $W_n(s)$ obtained in \cite{bnsw-rw}
term by term, we obtain
\begin{equation}\label{eq:WD0v2}
W_n'(0) = \log (n) - \sum_{m=1}^\infty \frac{1}{2m}
\sum_{k=0}^m \binom{m}{k} \frac{(-1)^k W_n(2k)}{n^{2k}}.
\end{equation}
On the other hand, from \cite{rtv-mahler} we find the strikingly similar
\begin{equation}\label{eq:WD0v}
W_n'(0) = \frac12 \log (n) - \frac{\gamma}{2} - \sum_{m=2}^\infty \frac{1}{2m}
\sum_{k=0}^m \binom{m}{k} \frac{(-1)^k W_n(2k)}{k! n^{k}}.
\end{equation}
Finally, we note that $W_n(s)$ itself is a special case of \textit{zeta Mahler
measure} as introduced recently in \cite{aka}.
\section{New results on the moments $W_n$}\label{sec:Wn}
From \cite{bbbg-bessel} equation (23), we have for $k>0$ even,
\begin{equation}\label{k3}
W_3(k) = \frac{3^{k+3/2}}{\pi \, 2^k\, \Gamma(k/2+1)^2} \int_0^\infty t^{k+1}K_0(t)^2 I_0(t) \mathrm{d} t,
\end{equation}
where $I_0(t), K_0(t)$ denote the \textit{modified Bessel functions} of the first and second kind, respectively.
Similarly, \cite{bbbg-bessel} equation (55) states that for $k>0$ even,
\begin{equation}\label{k4}
W_4(k) = \frac{4^{k+2}}{\pi^2 \, \Gamma(k/2+1)^2} \int_0^\infty t^{k+1}K_0(t)^3 I_0(t) \mathrm{d} t.
\end{equation}
Equation \eqref{k3} can be formally reduced to a closed form as a $_3F_2$ (for
instance by \textit{Mathematica}). At $k=\pm 1$, the closed form agrees
with $W_3(\pm 1)$. As both sides of \eqref{k3} satisfy the same recursion
(\cite{bbbg-bessel} equation (8)), we see that it in fact holds for all
integers $k>-2$.
In the following we shall use Carlson's theorem (\cite{titchmarsh}) which states:
\emph{Let
$f$ be analytic in the right half-plane $\Re z \ge 0$ and of exponential type
with the additional requirement that $$|f(z)| \le M e^{d |z|}$$ for some $d <
\pi$ on the imaginary axis $\Re z = 0$. If $f(k)=0$ for $k=0,1,2,\ldots$ then
$f(z)=0$ identically}.
We then have the following:
\begin{lemma} Equation (\ref{k3}) holds for all $k$ with $\mathrm{Re}\, k>-2$.
\end{lemma}
\begin{proof}
Both sides of (\ref{k3}) are of exponential type and agree when $k=0,1,2, \ldots$. The standard estimate shows that the right-hand side grows like $e^{|y| \pi/2}$ on the imaginary axis. Therefore the conditions of Carlson's theorem are satisfied and the identity holds whenever the right-hand side converges.
\end{proof}
Using the closed form given by the computer algebra system, we thus have:
\begin{theorem}[Single hypergeometric for $W_3(s)$]
For $s$ not a negative integer $<-1$,
\begin{equation}\label{singleh}
W_3(s) = \frac{3^{s+3/2}}{2\pi}\frac{\Gamma(1+s/2)^2}{\Gamma(s+2)}\,
\pFq32{\frac{s+2}2, \frac{s+2}2,\frac{s+2}2}{1,\frac{s+3}2}{\frac14}.
\end{equation}
\end{theorem}
Turning our attention to negative integers, we have for $k \ge 0$ an integer:
\begin{equation} \label{kn}
W_3(-2k-1) = \frac{4}{\pi^3} \left(\frac{2^k k!}{(2k)!}\right)^2 \int_0^\infty t^{2k}K_0(t)^3 \mathrm{d} t,
\end{equation}
because the two sides satisfy the same recursion (\cite[(8)]{bbbg-bessel}), and agree
when $k=0, 1$ (\cite[(47) and (48)]{bbbg-bessel}).
\begin{remark}
Equation (\ref{kn}) however does not hold when $k$ is not an integer. Also, combining (\ref{kn}) and (\ref{k3}) for $W_3(-1)$, we deduce
\[ \int_0^\infty K_0(t)^2 I_0(t)\, \mathrm{d} t = \frac{2}{\sqrt3 \pi} \int_0^\infty K_0(t)^3 \,\mathrm{d} t = \frac{\pi^2}{2\sqrt{3}} \int_0^\infty J_0(t)^3\, \mathrm{d} t. \]
\end{remark}
From (\ref{kn}), we experimentally determined a single hypergeometric for $W_3(s)$ at negative odd integers:
\begin{lemma} For $k \ge 0$ an integer,
\[ W_3(-2k-1) = \frac{\sqrt{3} \, \binom{2k}{k}^2}{2^{4k+1}3^{2k}} \, _3F_2\left( {{\frac12, \frac12, \frac12}\atop{k+1, k+1}}\bigg| \frac14 \right). \]
\end{lemma}
\begin{proof} It is easy to check that both sides agree at $k=0, 1$. Therefore we need only to show that they satisfy the same recursion. The recursion for the left-hand side implies a contiguous relation for the right-hand side, which can be verified by extracting the summand and applying Gosper's algorithm (\cite{aeqb}).
\end{proof}
The integral in (\ref{kn}) shows that $W_3(-2k-1)$ decays to 0 rapidly -- very roughly like $9^{-k}$ as $k \to \infty$ -- and so is never $0$ when $k$ is an integer.
To show that (\ref{k4}) holds for more general $k$ required more work. Using Nicholson's integral representation in \cite{watson-bessel}, \[I_0(t)K_0(t) = \frac{2}{\pi} \int_0^{\pi/2} K_0(2t \sin a) \,\mathrm{d} a,\]
the integral in (\ref{k4}) simplifies to
\begin{equation}\label{kk}
\frac{2}{\pi} \int_0^{\pi/2} \int_0^\infty t^{k+1} K_0(t)^2 K_0(2t \sin a) \,\mathrm{d} t \mathrm{d} a.
\end{equation}
The inner integral in (\ref{kk}) simplifies in terms of a \emph{Meijer G-function}; \textit{Mathematica} is able to produce
\[ \frac{\sqrt{\pi}}{8 \sin^{k+2} a}\,G_{3,3}^{3,2} \left( {{-\frac12, -\frac12, \frac12}\atop{0,0,0}}\bigg| \frac{1}{\sin^2 a}\right), \]
which transforms to
\[ \frac{\sqrt{\pi}}{8\sin^{k+2} a}\,G_{3,3}^{2,3} \left( {{1,1,1}\atop{\frac32,\frac32,\frac12}}\bigg| \sin^2 a\right). \]
Let $t = \sin^2 a$ in the above, so the outer integral in (\ref{kk}) transforms to
\begin{equation}\label{preeuler}
\frac{\sqrt{\pi}}{16} \int_0^1 t^{-\frac{k+3}2}(1-t)^{-\frac12}\, G_{3,3}^{2,3} \left( {{1,1,1}\atop{\frac32,\frac32,\frac12}}\bigg| t \right)
\,\mathrm{d} t.
\end{equation}
We can resolve this integral by applying the Euler-type integral
\begin{equation}\label{euler}
\int_0^1 t^{-a}(1-t)^{a-b-1} \, G_{p,q}^{m,n} \left({\mathbf{c}\atop\mathbf{d}} \bigg| z t \right) \mathrm{d} t = \Gamma(a-b)\, G_{p+1,q+1}^{m,n+1} \left({{a,\mathbf{c}}\atop{\mathbf{d},b}} \bigg| z \right).
\end{equation}
Indeed, when $k=-1$, the application of (\ref{euler}) recovers the Meijer G representation of $W_4(-1)$ (\cite{bsw-rw2}); that is, (\ref{k4}) holds for $k=-1$.
When $k=1$, the resulting Meijer G-function is
\[ G_{4,4}^{2,4} \left( {{2,1,1,1}\atop{\frac32,\frac32,\frac12,\frac32}}\bigg| 1 \right), \]
to which we apply Nesterenko's theorem (\cite{nest}), deducing a triple integral (up to a constant factor) for it:
\[ \int_0^1\int_0^1\int_0^1 \sqrt{\frac{x(1-x)z}{y(1-y)(1-z)(1-x(1-y z))^3}}\,\mathrm{d} x \mathrm{d} y\mathrm{d} z. \]
We can reduce the triple integral to a single integral,
\[ \int_0^1 \frac{8 E'(t)\big((1+t^2)K'(t)-2E'(t)\big)}{(1-t^2)^2} \,\mathrm{d} t. \]
Now applying the change of variable $t \mapsto (1-t)/(1+t)$, followed by quadratic transformations for $K$ and $E$, we finally get
\[ \int_0^1 \frac{4(1+t)}{t^2}E\left(\frac{2\sqrt{t}}{1+t}\right)\big(K(t)-E(t)\big) \,\mathrm{d} t, \]
which is, indeed, (a correct constant multiple times) the expression for $W_4(1)$ which follows from Section 3.1 in \cite{bsw-rw2}.
We finally observe that both sides of (\ref{k4}) satisfy the same recursion (\cite{bbbg-bessel} equation (9)), hence they agree for $k=0, 1, 2, \ldots$. Carlson's theorem applies with the same growth on the imaginary axis as in (\ref{k3}) and we have proven the following:
\begin{lemma}\label{L4} Equation (\ref{k4}) holds for all $k$ with $\mathrm{Re}\, k>-2$.
\end{lemma}
\begin{theorem}[Alternative Meijer G representation for $W_4(s)$] For all $s$,
\begin{equation}
W_4(s) = \frac{2^{2s+1}}{\pi^2 \, \Gamma(\frac12(s+2))^2} \, G_{4,4}^{2,4}\left( {{1,1,1,\frac{s+3}{2}}\atop{\frac{s+2}2,\frac{s+2}2,\frac{s+2}2,\frac12}}\bigg| 1\right).
\end{equation}
\end{theorem}
\begin{proof} Apply (\ref{euler}) to (\ref{preeuler}) for general $k$, and equality holds by Lemma \ref{L4}.
\end{proof}
Note that Lemma \ref{L4} combined with the known formula for $W_4(-1)$ in \cite{bsw-rw2} gives
\[ \frac{4}{\pi^3} \int_0^\infty K_0(t)^3I_0(t)\, \mathrm{d} t = \int_0^\infty J_0(t)^4\, \mathrm{d} t. \]
Armed with the knowledge of Lemma \ref{L4}, we may now resolve a very special but central case (corresponding to $n=2$) of Conjecture 1 in \cite{bsw-rw2}.
\begin{theorem}
For integer $s$,
\begin{equation} \label{conj}
W_4(s) = \sum_{j=0}^\infty \binom{s/2}{j}^2 W_3(s-2j).
\end{equation}
\end{theorem}
\begin{proof}
In \cite{bnsw-rw} it is shown that both sides satisfy the same three term recurrence, and agree when $s$ is even. Therefore, we only need to show that the identity holds for two consecutive odd values of $s$.
For $s=-1$, the right-hand side of (\ref{conj}) is
\[ \sum_{j=0}^\infty \binom{-1/2}{j}^2 W_3(-1-2j) = \sum_{j=0}^\infty \frac{2^{2-2j}}{\pi^3 j!^2} \int_0^\infty t^{2j} K_0(t)^3 \,\mathrm{d} t \]
upon using (\ref{kn}), and after interchanging summation and integration (which is justified as all terms are positive), this reduces to
\[ \frac{4}{\pi^3} \int_0^\infty K_0(t)^3I_0(t)\, \mathrm{d} t, \]
which is the value for $W_4(-1)$ by Lemma \ref{L4}.
We note that the recursion for $W_4(s)$ gives the pleasing reflection property
\[ W_4(-2k-1) \, 2^{6k} = W_4(2k-1). \]
In particular, $W_4(-3) = \frac{1}{64} \, W_4(1)$. Now computing the right-hand side of (\ref{conj}) at $s=-3$, and interchanging summation and integration as before, we obtain
\[ \sum_{j=0}^\infty \binom{-3/2}{j}^2 W_3(-3-2j) = \frac{4}{\pi^3} \int_0^\infty t^2 K_0(t)^3 I_0(t) \,\mathrm{d} t = \frac{1}{64}W_4(1) = W_4(-3). \]
Therefore (\ref{conj}) holds when $s=-1, -3$, and thus holds for all integer $s$.
\end{proof}
\bigskip
\paragraph{\textbf{Acknowledgments}}
We are grateful to David Bailey for significant numerical assistance, Michael
Mossinghoff for pointing us to the Mahler measure conjectures via
\cite{rtv-mahler}, and Plamen Djakov and Boris Mityagin for correspondence
related to Theorem \ref{thm:deleadingcoeffx} and the history of their proof.
We are specially grateful to Don Zagier for not only providing us with
his proof of Theorem \ref{thm:deleadingcoeffx} but also for his enormous
amount of helpful comments and improvements.
We also thank the reviewer for helpful suggestions.
The first and the last authors thankfully acknowledge support by the Australian
Research Council.
\section*{Appendix. \textbf{A family of combinatorial identities}}
\medskip
\hbox to\hsize{\hss \uppercase{Don Zagier}\footnote{The original note is unchanged.}\hss}
\vskip5mm
The ``collateral result'' of Djakov and Mityagin, \cite{dm1, dm2}, is the pair
of identities
\begin{align*}
&\sum_{\substack{ -m<i_1<\dots<i_k<m\\ i_2-i_1,\dots,i_k-i_{k-1}\ge2 }}
\;\,\prod_{s=1}^k(m^2-i_s^2)
= \sigma_k(1^2,3^2,\dots,(2m-1)^2)\,,
\displaybreak[2]\\
&\sum_{\substack{ 1-m<i_1<\dots<i_k<m\\ i_2-i_1,\dots,i_k-i_{k-1}\ge2 }}
\;\,\prod_{s=1}^k(m-i_s)(m+i_s-1)
= \sigma_k(2^2,4^2,\dots,(2m-2)^2)\,,
\end{align*}
where $m$ and $k$ are integers with $m\ge k\ge0$ and $\sigma_k$ denotes the $k$th elementary symmetric
function. By setting $j_s=i_s+m$ in the first sum and
$j_s=i_s+m-1$ in the second, we can rewrite these formulas more uniformly as\footnote{Note that \eqref{eq:FMk} is precisely Theorem \ref{thm:deleadingcoeffx}.}
\begin{align}\label{eq:FMk}
F_{M,k}(M) = \begin{cases}
\sigma_k(1^2,3^2,\dots,(M-1)^2)&\text{if $M$ is even,} \\
\sigma_k(2^2,4^2,\dots,(M-1)^2)&\text{if $M$ is odd,}
\end{cases}
\end{align}
where $F_{M,k}(X)$ is the polynomial in $X$ (non-zero only if $M\ge2k\ge0$) defined by
\begin{align}\label{eq:FMkX}
F_{M,k}(X) = \sum_{\substack{ 0<j_1<\dots<j_k<M\\ j_2-j_1,\dots,j_k-j_{k-1}\ge2 }}
\prod_{s=1}^k j_s(X-j_s)\;.
\end{align}
The advantage of introducing the free variable $X$ in \eqref{eq:FMkX} is that
the functions $F_{M,k}(X)$ satisfy the recursion
\begin{align}\label{eq:FMkrec}
F_{M+1,k+1}(X) - F_{M,k+1}(X) = M\,(X-M)\,F_{M-1,k}(X),
\end{align}
because the only paths that are counted on the left are those with $0<j_1<\dots<j_k<j_{k+1}=M$.
It is also advantageous to introduce the polynomial generating function
$$ \Phi_M = \Phi_M(X,u) = \sum_{0\le k\le M/2}(-1)^k\,F_{M,k}(X)\,u^{M-2k}\,, $$
the first examples being
\begin{align*}
&\Phi_0=1\,,\qquad \Phi_1=u\,,\qquad \Phi_2=u^2-(X-1)\,,\qquad \Phi_3=u^3-(3X-5)u\,,\\
&\Phi_4=u^4 -(6X-14)u^2 +(3X^2-12X+9)\,,\\
&\Phi_5=u^5-(10X-30)u^3+(15X^2-80X+89)\,,\\
&\Phi_6= u^6 - (15X - 55)u^4 + (45X^2 - 300X + 439)u^2 - (15X^3 - 135X^2 + 345X - 225)\,.
\end{align*}
In terms of this generating function, the recursion \eqref{eq:FMkrec} becomes
\begin{align}\label{eq:Phirec}
\Phi_{M+1} = u\,\Phi_M - M(X-M)\,\Phi_{M-1}
\end{align}
and the identity \eqref{eq:FMk} to be proved can be written succinctly as
\begin{align}\label{eq:idPhi}
\Phi_M(M,u) = \prod_{\substack{ |\lambda|<M\\ \lambda\not\equiv M\pmod2 }} (u-\lambda)\;.
\end{align}
Denote by $P_M(u)$ the polynomial on the right-hand side of \eqref{eq:idPhi}. Looking for other pairs $(M,X)$
where $\Phi_M(X,u)$ has many integer roots, we find experimentally that this happens whenever $M-X$
is a non-negative integer, and studying the data more closely we are to conjecture the two formulas
\begin{align}\label{eq:Phi1}
\Phi_M(M-n,u) = \frac1{2^n}\,\sum_{j=0}^n\binom nj\,P_M(u-n+2j)\qquad(M,\,n\ge0)
\end{align}
(a generalization of \eqref{eq:idPhi}) and
\begin{align}\label{eq:Phi2}
\Phi_{M+n}(M,u) = \Phi_M(M,u)\,\Phi_n(-M,u)\qquad(M,\,n\ge0)\,.
\end{align}
Formula \eqref{eq:Phi2} is easy to prove, since it holds for $n=0$ trivially and for $n=1$ by~\eqref{eq:Phirec}
and since both sides satisfy the recursion $y_{n+1}=u\,y_n+n(M+n)\,y_{n-1}$ for $n=1,2,\dots$ by \eqref{eq:Phirec}.
On the other hand, combining \eqref{eq:idPhi}, \eqref{eq:Phi1} and \eqref{eq:Phi2} leads to the conjectural formula
$$ \Phi_n(-M,u) = \frac1{2^n}\,\sum_{j=0}^n\binom nj\,\frac{P_{M+n}(u-n+2j)}{P_n(u)}
= n!\,\sum_{j=0}^n(-1)^j\binom{\frac{-u-M-1}2}j\,\binom{\frac{u-M-1}2}{n-j} $$
or, renaming the variables,
\begin{align}\label{eq:Phi3}
\frac1{M!}\,\Phi_M(x+y+1,y-x) = \sum_{j=0}^M(-1)^j\binom xj\,\binom y{M-j}\;.
\end{align}
To prove this, we see by \eqref{eq:Phirec} that, denoting by $G_M=G_M(x,y)$ the expression on the right,
it suffices to prove the recursion $(M+1)G_{M+1}=(y-x)G_M+(M-x-y-1)G_{M-1}$. This is an easy binomial
coefficient identity, but once again it is easier to work with generating functions: the sum
\begin{align}\label{eq:sumG}
\mathcal G(x,y;T) := \sum_{M=0}^\infty G_M(x,y) \,T^m = (1-T)^x\,(1+T)^y
\end{align}
satisfies the differential equation
$$\dfrac1{\mathcal G}\dfrac{\partial\mathcal G}{\partial T}=\dfrac y{1+T}-\dfrac x{1-T}$$
or
$$\dfrac{\partial\mathcal G}{\partial T}=(y-x)\,\mathcal G + \left(T\dfrac\partial{\partial T}-x-y\right)\mathcal G,$$
and this is equivalent to the desired recursion.
We can now complete the proof of \eqref{eq:FMk}. Rewriting \eqref{eq:sumG} in the form
$$ \frac1{M!}\,\Phi_M(X,u) = \text{coeff}_{T^M}\left((1-T)^{\frac{X-u-1}2}\,(1+T)^{\frac{X+u-1}2}\right)\,, $$
we find that, for $1\le j\le M$,
$$\frac1{M!}\, \Phi_M(M,M+1-2j) = \text{coeff}_{T^M}\left((1-T)^{j-1}\,(1+T)^{M-j}\right) = 0 $$
and hence that the polynomial on the left-hand side of \eqref{eq:idPhi} is divisible by the polynomial on the right,
which completes the proof since both are monic of degree~$M$ in~$u$.
|
\subsection{X-ray constraints}
In steady nuclear burning regime, corresponding to the mass accretion rates $\dot{M}\ge 10^{-7}~M_\odot$/yr, energy of hydrogen fusion is liberated in the form of electromagnetic radiation, with luminosity of
$$L_{nuc}=\epsilon_H X \dot{M} \sim 10^{37} {\rm ~ erg/s}$$
where $\epsilon_H\approx 6\cdot 10^{18}$ erg/g is energy release per unit mass and $X$ -- hydrogen mass fraction (the solar value of $X=0.72$ is assumed).
The nuclear luminosity exceeds by more than an order of magnitude the gravitational energy of accretion, $L_{grav}=GM\dot{M}/R$, and maintains the effective temperature of the WD surface at the level, defined by the Stefan-Boltzmann law:
\begin{equation}
T_{eff}\approx 67
\left(\frac{\dot{M}}{5\cdot 10^{-7} M_\odot/yr}\right)^{1/4}
\left(\frac{R_{WD}}{10^{-2}R_\odot}\right)^{-1/2}
{\rm~eV}
\end{equation}
Such a soft spectrum is prone to absorption by interstellar gas and dust, especially at smaller temperatures. Because the WD radius $R_{WD}$ decreases with its mass \cite{panei}, the $T_{eff}$ increases as the WD approaches the Chandrasekhar limit -- the signal, detectable at X-ray wavelengths, will be dominated by the most massive WDs (Fig.\ref{fig:kt}) \cite{nat2010, distefano2010apj}.
The type Ia supernova rate scales with near-infrared luminosity of the host galaxy and for E/S0 galaxies $\dot{N}_{SNIa}\approx 3.5\cdot 10^{-4}$ yr$^{-1}$ per $10^{10}~L_{K,\odot}$ \cite{mannucci}.
If the WD mass grows at a rate $\dot{M}$, a population of
\begin{equation}
N_{WD}\sim \frac{\Delta M}{\dot{M}\left<\Delta t\right>}\sim \frac{\Delta M}{\dot{M}}\dot{N}_{SNIa}
\label{eq:nwd}
\end{equation}
accreting WDs is needed in order for one supernova to explode on average every $\left<\Delta t\right>=\dot{N}_{SNIa}^{-1}$ years. Thus, for a typical galaxy, the accretion scenario predicts a numerous population of accreting white dwarfs, $N_{WD}\sim {\rm few} \times (10^2 - 10^3)$, much more than numbers of soft X-ray sources actually observed \cite{distefano2010an}.
Therefore, although the brightest and hottest of them may indeed reveal themselves as super-soft sources, the vast majority of SN Ia progenitors must remain unresolved or hidden from the observer, for example by interstellar absorption, in order for the accretion scenario to work.
Their combined luminosity is
\begin{equation}
L_{tot}=L_{nuc}\times N_{WD}=\epsilon X\Delta M\dot{N}_{SNIa}
\label{eq:ltot}
\end{equation}
where $\Delta M$ is the difference between the Chandrasekhar mass and the initial WD mass. Predicted luminosity can be compared with observations, after absorption and bolometric corrections are accounted for.
\begin{table}
\caption{Comparison of the accretion scenario with observations \cite{nat2010}.
X-ray luminosities refer to the soft (0.3--0.7 keV) band. The columns marked "predicted" display total number and combined X-ray luminosity (absorption applied) of accreting WDs in the galaxy predicted in case the single degenerate scenario would produce all SNe Ia. In computing predicted numbers we halved the SN Ia rates as discussed in \cite{nat2010}, other parameters are: $\dot{M}=10^{-7}~M_\odot$/yr, initial WD mass $1.2M_\odot$. }
\vspace{0.5cm}
\centering
\begin{tabular}{l c c c c}
\hline
Name~~~~~~~~~ & $L_K$ [$L_{K,\odot}$] & ~~~$N_{WD}$~~~ &\multicolumn{2}{c}{$L_X$ [erg/s]} \\
& ~~observed~~ & ~~predicted~~ &~~observed~~ & ~~predicted~~ \\
\hline
M32 & $8.5\cdot 10^{ 8}$ & $25$ &$1.5\cdot 10^{36}$ & $7.1\cdot 10^{37}$ \\
NGC3377 & $2.0\cdot 10^{10}$ & $5.8\cdot 10^2$ &$4.7\cdot 10^{37}$ & $2.7\cdot 10^{39}$ \\
M31 bulge & $3.7\cdot 10^{10}$ & $1.1\cdot 10^3$ &$6.3\cdot 10^{37}$ & $2.3\cdot 10^{39}$ \\
M105 & $4.1\cdot 10^{10}$ & $1.2\cdot 10^3$&$8.3\cdot 10^{37}$ & $5.5\cdot 10^{39}$ \\
NGC4278 & $5.5\cdot 10^{10}$ & $1.6\cdot 10^3$ &$1.5\cdot 10^{38}$ & $7.6\cdot 10^{39}$ \\
NGC3585 & $1.5\cdot 10^{11}$ & $4.4\cdot 10^3$ &$3.8\cdot 10^{38}$ & $1.4\cdot 10^{40}$ \\
\hline
\end{tabular}
\label{tab:lx}
\end{table}
To this end we collected archival data of X-ray (Chandra) and near-infrared (Spitzer and 2MASS) observations of several nearby gas-poor elliptical galaxies and for the bulge of M31 (\cite{bogdan2010,nat2010}, Table \ref{tab:lx}). Using K-band measurements to predict the SN Ia rates, we computed combined X-ray luminosities of SN Ia progenitors expected in the accretion scenario and compared them with Chandra observations.
Obviously, the observed values present upper limits on the luminosity of the hypothetical population of accreting WD, as there may be other types of X-ray sources contributing to the observed emission.
As is clear from the Table \ref{tab:lx}, predicted luminosities surpass observed ones by a factor of $\sim 30-50$, i.e. the accretion scenario is inconsistent with observations by a large margin.
\subsection{Statistics of recurrent and classical novae}
Unstable nuclear burning at low $\dot{M}$ is not generally believed to allow accumulation of mass, sufficient to lead to supernova explosion. However, just below the stable burning limit a considerable fraction of the envelope mass can be retained by the WD during the nova explosion \cite{prialnik}.
This motivated some authors to propose recurrent novae as supernova progenitors \cite[e.g.][]{hachisu2001}.
We demonstrate below that such systems will overproduce nova explosions in galaxies.
Assuming that classical/recurrent nova sources are the main type of SN Ia progenitors,
one can relate the nova and supernova rates:
\begin{equation}
\Delta M_{acc}\, \dot{N}_{CN} \sim \Delta M_{SN}\, \dot{N}_{SN}
\end{equation}
where $\Delta M_{acc}\le 10^{-7}-10^{-4}$ M$_\odot$ is the mass accumulated by the WD per one nova outburst cycle,
$\Delta M_{SN}\sim 0.5$ M$_\odot$ is the mass needed for the WD to reach the Chandrasekhar limit.
As $\Delta M_{acc}$ depends on the $\dot{M}$ and WD mass (Fig.\ref{fig:cn}), we write more precisely:
\begin{equation}
\frac{\dot{N}_{CN}}{\dot{N}_{SNIa}}=\int\frac{dM_{WD}}{\Delta M_{acc}(M_{WD},\dot{M})}\ge
\int\frac{dM_{WD}}{\Delta M_{CN}(M_{WD},\dot{M})}
\label{eq:cn2sn}
\end{equation}
where $\Delta M_{CN}$ is the mass of the hydrogen shell required for the nova outburst to start. The inequality in the eq.(\ref{eq:cn2sn}) follows from the fact that $\Delta M_{acc}\le \Delta M_{CN}$, due to the possible mass loss during the nova outburst.
As the $\Delta M_{CN}$ decreases steeply with the WD mass (Fig.\ref{fig:cn}), the main contribution to the predicted CN rate is made by the most massive WDs, similar to X-ray emission in the steady nuclear burning regime. They will be producing frequent outbursts with relatively short decay times \cite{prialnik}, thus resulting in a large population of fast (some of them recurrent) novae. This will contradict to the statistics of CNe, as illustrated by the example of M31 shown in Fig.\ref{fig:cn}.
Indeed, the observed rate of CN with decay time shorter than 20 days in this galaxy is $\approx 5.2\pm1.1$ yr$^{-1}$ \cite{capaccioli}, while eq.(\ref{eq:cn2sn}) predicts $\sim 300$ yr$^{-1}$ for the mass accretion rates relevant to the recurrent novae-based progenitors models, $\dot{M}\sim10^{-8}$ M$_\odot$/yr. As $\Delta M_{CN}$ is larger at small $\dot{M}$ (Fig.\ref{fig:cn}), the contradiction between observed and predicted nova frequencies becomes less dramatic at smaller $\dot{M}$. However, very low values of $\dot{M}\ll 10^{-10}$ M$_\odot$/yr are not feasible in the context of SN Ia progenitors.
More realistic models with $\dot{M}\ge 10^{-8}$ M$_\odot$/yr do not produce more that $\sim 2\%$ of type Ia supernovae.
\begin{figure}
\centering
\hbox{
\includegraphics[height=.35\textheight]{cn_mwd.ps}
\hspace{0.2cm}
\includegraphics[height=.30\textheight]{cn_hist.ps}
}
\caption{{\em Left:} The $\dot{M}$ and $M_{WD}$ dependence of the nova ignition mass $\Delta M_{CN}$ and decay time
$t_3$, from \cite{prialnik}. {\em Right:} The decay time $t_3$ distribution for CNe in M31. The histogram shows the observed distribution from \cite{capaccioli} normalized to the total rate of 25 CN/yr. The systematic effects may somewhat modify its shape but will not affect our conclusions, as the observations control times do not vary significantly across the $t_3$ range of interest \cite{capaccioli}.
Two smooth lines show predicted distributions for two different values of $\dot{M}$. They are computed assuming that cataclysmic variables are the sole progenitors of SN Ia and are based on the results of numerical simulations of CN outburst cycles from \cite{prialnik} for the WD core temperature of $10^7$ K. The vertical dashed line denotes the boundary of the fast Novae, which statistics is discussed in the text.
}
\label{fig:cn}
\end{figure}
\subsection{Conclusion}
Thus, no more than $\sim$few per cent of SNe Ia in early type galaxies can be produced by\ white dwarfs accreting from a donor star in a binary system and exploding at the Chandrasekhar mass limit.
In the steady nuclear burning regime the supernova progenitors would emit too much of soft X-ray emission, while if nuclear burning is unstable they would overproduce classical nova explosions.
At very high $\dot{M}$ the white dwarf could grow in mass without conflicting X-ray constraints or nova statistics, but would do this rather inefficiently, because a significant fraction of the transferred mass is lost in the wind \cite{hachisu, vdh97}. Therefore a relatively massive ({\em at least} $M\ge 1.3-1.7$ M$_\odot$) donor star is required for the white dwarf to reach the Chandrasekhar limit in this regime. Because the lifetimes of such stars do not exceed $\sim$few Gyr, this mechanism may work only in late-type and in the youngest of early-type galaxies.
As relevance of sub-Chandraskhar models is still debated \cite{hoeflich96, nugent97, sim2010}, the only currently viable alternative are WDs mergers \cite{iben, webbink}. This mechanism may be the main formation channel for SN Ia in early type galaxies. In late-type galaxies, on the contrary, massive donor stars are available, making the mass budget less prohibitive, so that WDs can grow to the Chandrasekhar mass entirely inside an optically thick wind or, via accretion of He-rich material from a He donor star \cite{iben94}. In addition, a star-forming environment is usually characterized by large amounts of neutral gas and dust, leading to increased absorption obscuring soft X-ray radiation from accreting WDs.
Therefore in late-type galaxies the accretion scenario may play a significant role, explaining, for example, the population of prompt supernovae \cite{bogdan2010a}.
\bibliographystyle{aipproc}
|
\section{Introduction}
One of the leading candidates for the mechanism responsible for the acceleration of Galactic cosmic rays is diffusive shock acceleration (DSA).
Proposed originally by Krimsky (1977), Axford, Leer \& Skadron (1977), Bell (1978a,b) and Blandford \& Ostriker (1978), the theory is now
well-understood in the case of non-relativistic shocks (shock speed $V_{\rm s} \ll c$), see for instance the reviews of Drury (1983),
Blandford \& Eichler (1987), Achterberg (2001) and Malkov \& Drury (2001). In the test-particle limit in a steady flow, where the accelerated particles have no influence on the flow,
DSA at a strong, infinitely thin hydrodynamic shock yields a power-law distribution in momentum for the accelerated particles with a spectral
index $q = - {\rm d} \ln n(p)/{\rm d} \ln p \simeq 2$, where $p$
is the particle momentum and $n(p)$ the momentum distribution.
Such a power law is inferred for cosmic rays at the source between several GeV/nucleon and $\sim 100$ TeV per nucleon, after a correction of the
spectrum observed at Earth for the effects of propagation inside (and escape from) the Galaxy.
In time-dependent flows with a complex flow geometry, in flows that contain multiple shocks and in particular when the cosmic rays through their pressure
significantly decelerate the pre-shock flow, analytical results are more difficult to obtain. In those cases one often has to resort to numerical methods to solve the basic equations.
When simulating DSA
one essentially tries to determine the cosmic ray density $N(\bm{x} \: , \: p \: , \: t)$
in reduced phase space $(\bm{x} \: , \: p = |\bm{p}|)$ as a
function of time. Two main approaches are possible: one either solves the Vlasov equation for the distribution function $N(\bm{x} \: , \: p \: , t)$
directly, or one uses an algorithm that constructs $N(\bm{x} \: , \: p \: , \: t)$ from particle positions in phase space that have been obtained
by direct simulation of representative particle orbits. The first approach requires the solution of a partial differential equation in (reduced) phase space, which is
generally computationally expensive as it requires matrix inversion in schemes such as the Crank-Nicholson method (e.g. Potter, 1973) that is needed
in order to stably solve a diffusion-advection type equation with sufficient accuracy. The advantage of course is that one obtains the distribution function directly.
The second approach is simple to implement in arbitrary geometries and can use very accurate integration schemes as one in principle solves an ordinary differential equation.
Its disadvantage is that one must take measures such as particle splitting (see below) in order to minimize the effect of Poisson noise that is unavoidable as one uses
a finite number of particles to construct the distribution function.
This paper is concerned with the second method, which involves the solution of {\em stochastic differential equations} (SDEs).
It was first proposed in this context by Achterberg \& Kr\"ulls (1992). Recent applications of this method include Marcowith \& Casse (2010) and Schure {\em et al,} (2010).
We describe a relatively simple stochastic predictor-corrector scheme for diffusive shock acceleration. It works well
in the presence of strong gradients in the diffusivity of the particles. In that case this scheme
is considerably more accurate than the simple Cauchy-Euler scheme,
which fails to return accurate results for the spectral slope $q$. The slope returned by the Cauchy-Euler scheme is consistently too steep, which indicates that
the shock transition (where the acceleration takes place) is not sampled accurately.
If the diffusivity gradient is small or vanishes, both schemes give almost identical results.
We outline the different numerical approaches in Section 2, discuss the need for a better scheme in Section 3, introduce a predictor-corrector type scheme in Section 4 and
evaluate its performance in Section 5. Conclusions are found in Section 6.
\section{Numerical approaches to DSA using stochastic differential equations}
The aim of this paper is to construct the energy distribution of particles (cosmic rays) that are accelerated in a prescribed flow containing a shock,
thereby limiting ourselves to the test-particle case.
The particles undergo spatial diffusion with respect to the flow
as a result of pitch angle scattering on magnetic fluctuations, and gain (or lose) energy due to compression (expansion) of the fluid.
Other processes, such as radiation losses and stochastic acceleration by plasma waves, can be added in a simple manner but will be neglected here.
We consider the simplest case of a one-dimensional, steady flow along the $x$-axis with fluid velocity $V(x)$.
Let
\be
\label{Fdef}
N(x \: , \: y \: , \: t) \equiv \frac{{\rm d} {\cal N}}{{\rm d}x \: {\rm d}y}
\ee
be the cosmic ray distribution function, where $y$ is a logarithmic momentum variable, defined formally as
\be
\label{ydef}
y = \ln (p/mc) \; .
\ee
Here $m$ is the particle rest mass and $c$ is the velocity of light. The particles are coupled to the fluid by frequent pitch angle scattering
by scattering centers (magnetic field fluctuations due to hydromagnetic waves) that are themselves tied to the flow. In the simplest case,
where these field fluctuations are advected passively by the flow with the local flow velocity $V(x)$,
the cosmic ray distribution $N(x \: , \: y \: , \: t)$
satisfies the equation (cf. Skilling, 1975):
\be
\label{Skeqn}
\frac{\partial N}{\partial t} + \frac{\partial}{\partial x} \left( V(x) \: N - D(x) \: \frac{\partial N}{\partial x} \right) =
\frac{1}{3} \left( \frac{{\rm d} V}{{\rm d} x} \right) \: \frac{\partial N}{\partial y} \; .
\ee
Here $N(x \: , \: y \: , \: t)$ is the distribution (averaged over gyration phase in the laboratory frame,
the shock rest frame in the application discussed below.
$D(x)$ is the position-dependent spatial diffusion coefficient.
For simplicity we assume this quantity to be independent of $p$ (and $y$), but this is
not important for what follows. The limitation to a steady flow is also not fundamental: the same techniques can be used to follow cosmic rays
in a time-varying flow.
It is known that the solutions to this Fokker-Planck type equation, which essentially expresses the propagation
of particles in the two-dimensional phase space $(x \: , \: y)$, can be constructed by following particles that satisfy a stochastic differential equation
(SDE) (e.g. Gardiner, 1983, Ch.5; {\O}ksendal, 1992, Ch. 3).
\subsection{Spatial transport}
Spatial transport in this case is described by a SDE of the type (Achterberg \& Kr\"ulls, 1992):
\be
\label{SDE}
{\rm d}x = U(x ) \: {\rm d}t + \sqrt{2 D(x)} \: {\rm d} W_{t} \; .
\ee
In Eqn. (\ref{SDE}) the quantity
\be
\label{Udef}
U(x) \equiv V(x) + \frac{{\rm d} D(x)}{{\rm d} x}
\ee
is an effective advection velocity that includes a drift term due to the
dynamical friction that results from a spatial gradient in the diffusivity.
The first term in
SDE (\ref{SDE}) is a deterministic term $\propto {\rm d} t$ that describes the average flow of particles
while the second stochastic term represents the spatial diffusion.
That last term involves a infinitesimal Wiener process ${\rm d}W_{t}$, which satisfies
\be
\left< {\rm d} W_{t} \right> = 0 \; \; \; , \; \; \; \left< {\rm d} W_{t}^2 \right> = {\rm d}t \; .
\ee
The brackets represent an average over many statistically independent realizations of this Wiener process.
In practice, this average is achieved
by simulating a large number of particles with statistically independent orbits.
The distribution of these particles in phase space approaches the solution of the corresponding Fokker-Planck equation,
provided one uses a sufficient number of particles
in order to minimize the effects of Poisson noise.
In the context of Diffusive Shock Acceleration the typical length scale in the cosmic ray precursor ahead of the shock is the diffusion length
\be
L_{\rm diff} = D/V \; .
\ee
This implies that the drift velocity within the precursor is of order
\be
V_{\rm drift} = \frac{{\rm d} D(x)}{{\rm d} x} \simeq \frac{\Delta D}{L_{\rm diff}} = \left( \frac{\Delta D}{D} \right) \: V \; .
\ee
Taking $\Delta D \simeq D$ and $V \simeq V_{\rm s}$, with $V_{\rm s}$ the shock speed, one finds that $V_{\rm drift} \simeq V_{\rm s}$.
Formally problems arise if $V_{\rm drift} \simeq v$ with $v$ the particle velocity.
However, in that case the diffusion approximation that is the basis of this method fails.
The diffusion approximation presupposes that
the anisotropy in the {\em exact} momentum distribution of the accelerated particles, which is of order $V_{\rm s}/v$, is small.
In relativistic shocks, where $V_{\rm s} \simeq c$, this is never the case.
When $V_{\rm drift} \simeq v$ ($\simeq c$ for relativistic particles)
one has to simulate the pitch angle scattering of particles directly, as is routinely done in simulations of particle acceleration
at relativistic shocks (e.g. Bednarz \& Ostrowski, 1998; Achterberg {\em et al.}, 2001).
Within the shock transition itself one expects a change in diffusivity $\Delta D \simeq D$ over the shock width $L_{\rm s}$ so that
\be
V_{\rm drift} \simeq \frac{D}{L_{\rm s}} \simeq v \: \left( \frac{\lambda_{\rm mfp}}{3 L_{\rm s}} \right) \; .
\ee
Here $\lambda_{\rm mfp}$ is the scattering mean free path so that the diffusion coefficient equals $D = v \lambda_{\rm mfp}/3$.
The diffusion approximation requires $V_{\rm drift} \ll v$ so {\em formally} the method fails if $L_{\rm s} \le \lambda_{\rm mfp}$.
However, in that case one may as well approximate the shock as a discontinuity with a stepwise jump in the diffusivity and flow speed.
A scaling method for dealing with such sudden step-like jumps in the diffusion coefficient and velocity has been devised by Zhang (2000).
It can be applied in this case so that one solves an equivalent advection-diffusion problem where
{\em no} drift term due to the diffusivity jump at the shock occurs, and the method outlined below is valid as long as the diffusion approximation applies
to particles ahead of and behind the shock front.
\subsection{Change in particle momentum}
Transport in momentum (in the absence of radiation losses or second-order Fermi acceleration by waves) follows from
\be
\label{momch}
{\rm d}y \simeq \frac{{\rm d} p}{p} = - \frac{1}{3} \left( \frac{{\rm d} V}{{\rm d} x} \right) \: {\rm d}t \equiv \omega(x) \: {\rm d}t \; .
\ee
This gives the momentum changes in response to compressions or rarefactions in the flow, with $\omega(x)$ the local acceleration rate.
The effect of radiation losses and of second-order Fermi acceleration can be added in a simple manner.
\subsection{The Cauchy-Euler scheme}
The simplest numerical scheme that one can use to simulate this diffusion-advection process is the explicit
Cauchy-Euler scheme (CES), see for instance Achterberg \& Kr\"ulls, 1992. This scheme
advances $x$ and $y$ in time (time step: $\Delta t$) with increments in position and log momentum $\Delta x$ and $\Delta y$ given by:
\ba
\label{CES}
\Delta x & = & U(x) \: \Delta t + \sqrt{2 D(x) \Delta t} \: \xi_{t} \equiv \Delta x_{\rm s} + \Delta x_{\rm diff}
\: \xi_{t} \nonumber \\
& & \\
\Delta y & = & \omega(x) \: \Delta t \; . \nonumber
\ea
Here $\xi_{t}$ is a normally distributed Wiener process with zero average and unit dispersion, in the notation of Kloeden \& Platen (1992):
\be
\label{normdist}
\xi_{t} \in {\rm N}(0 , 1) \; .
\ee
The quantity
\be
\label{Dstep}
\Delta x_{\rm diff} \equiv
\sqrt{2 D(x) \Delta t}
\ee
is the rms diffusive step, given the time increment $\Delta t$. The statistically sharp (non-stochastic) step is (see Eqn. \ref{Udef})
\be
\Delta x_{\rm s} = V \: \Delta t + \frac{{\rm d} D}{{\rm d} x} \: \Delta t \equiv \Delta x_{\rm adv} + \Delta x_{\rm drift} \; .
\ee
It consists of the advective step $\Delta x_{\rm adv} = V \: \Delta t$ due to the plasma flow
and the drift term $\Delta x_{\rm drift} \propto {\rm d} D/{\rm d} x$ due to gradients in the diffusivity.
This scheme has the virtue of simplicity, and for small $\Delta x_{\rm drift}$ it can accurately describe cosmic ray acceleration near shocks with a
thickness $L_{\rm s}$ provided the time step is chosen
in such a way that
\be
\label{acccond}
\Delta x_{\rm adv} \ll L_{\rm s} \ll \Delta x_{\rm diff}
\ee
(Achterberg \& Kr\"ulls, 1992), who only tested the CES for uniform diffusivity so that $\Delta x_{\rm drift} = 0$.
The first condition is necessary to resolve the shock transition, and is essentially a demand of sufficient accuracy.
The second condition ensures that a test particle, while crossing the shock with average
velocity $U$ and before escaping into the downstream flow, crosses the shock many times in diffusive jumps that are typically a few times the shock thickness.
The simulated particles essentially mimic the behavior of a real cosmic ray undergoing acceleration near a strong shock, such as a supernova blast wave.
The whole procedure can be thought of as a replacement of the real diffusion (with a microscopic step size) by an equivalent diffusion process with the
{\em same} diffusion coefficient but with a macroscopic step size.
The last condition then ensures that the acceleration rate $\omega(x)$
due to the compression inside the shock transition
is sampled frequently and in a stochastic manner. In this way the simulation produces the correct spectrum of accelerated particles.
Marcowith \& Kirk (1999) have proposed a partially
implicit scheme that uses the new particle position $x + \Delta x$ to evaluate $\Delta y$.
They showed in the simple case of a linear shock transition and a {\em constant}
diffusivity $D$ (so that once again $\Delta x_{\rm drift} = 0$)
that this approach leads to good results, while the condition (\ref{acccond}) can be relaxed to $\Delta x_{\rm adv} \ll \Delta x_{\rm diff}$.
Their approach opens the possibility of using larger time steps and considering zero-thickness shocks, where the velocity change is modeled by a step function.
In our approach outlined below we will follow their method for the calculation of the momentum change in terms of $\Delta y$, but will need to
keep an explicit algorithm for the integration of the SDE for $\Delta x$.
\section{The need for a better scheme}
During numerical experiments, where acceleration of test particles is calculated in a flow that is obtained using a numerical MHD code,
it came to our attention that the simple Cauchy-Euler scheme
does not yield satisfactory results if the drift due to gradients in the diffusivity becomes too large, in particular when
$|\Delta x_{\rm drift}| \gg |\Delta x_{\rm adv}|$.
If the scale length of the variation in the diffusion coefficient is
$L_{\rm d} = |(1/D)(\partial D/\partial x)|^{-1}$ one has
\be
\label{dxratio}
\left| \frac{\Delta x_{\rm drift}}{\Delta x_{\rm adv}} \right| \simeq \frac{D}{L_{\rm d} V} \equiv \frac{L_{\rm diff}}{L_{\rm d}} \; .
\ee
Here $L_{\rm diff} = D/V$ is the characteristic length of DSA, the {\em diffusion length} introduced above.
We have found that the CES fails to give the correct particle distribution when
\be
L_{\rm d} \simeq L_{\rm s} \ll L_{\rm diff} = \frac{D}{V}\; .
\ee
In particular (as illustrated below) the momentum distribution of the accelerated particles produced by the CES for reasonable values of the time step $\Delta t$
is too steep. This implies that the average momentum gain per
shock crossing as calculated using the CES is too low and the acceleration rate $\omega(x)$ is apparently not sampled accurately by the particles following the
simulated orbits. We reiterate that
Achterberg \& Kr\"ulls (1992) and Marcowith \& Kirk (1999) did not consider this case in their calculations: they assumed a constant diffusivity $D$.
The case of a diffusivity that varies on a scale comparable with the shock transition is astrophysically important.
For instance: it is often assumed that the scattering
of cosmic rays near an accelerating shock is due to saturated Alfv\'en wave turbulence where the diffusivity scales as $D \propto B^{-1}$, the case of Bohm diffusion.
In a shock in an infinitely conducting plasma the MHD shock conditions imply that $B$ increases across the shock by a factor
\be
\label{rb}
r_{\rm B} = \sqrt{ \cos^2 \theta_{\rm B} + r^2 \: \sin^2 \theta_{\rm B}} \; .
\ee
The parameter $r = \rho_{2}/\rho_{1}$ is the shock compression ratio, and $\theta_{\rm B} = \cos^{-1}(\hatn \bdot \bm{B}_{1})$ is the inclination angle
between the upstream magnetic field $\bm{B}_{1}$ and the normal $\hatn$ to the shock surface. Depending on the orientation of the upstream magnetic field
one has $1 \le r_{\rm B} \le r$. Additional changes in the diffusivity can arise through the reflection and transmission of MHD waves at the shock.
Here (and in what follows) we will use the subscripts 1 (2) to describe the value of quantities ahead of (behind) the shock transition. A similar case is obtained if
the cosmic rays lead to field amplification through the Bell-Lucek instability (Bell \& Lucek, 2001): there the field is amplified on the diffusive scale $L_{\rm diff} \sim D/V$,
in addition to the amplification by compression at the shock.
The use of a fully implicit scheme is not feasible for our application, as the velocity field $V(x \: , \: t)$ and the magnetic field $B(x \: , \: t)$
are not analytic functions but are determined by a MHD code.
We therefore need a scheme that is explicit for the advance of the position $x$, that is: it uses the variables at time $t$ and old position $x$
to calculate the change $\Delta x$ and the new position $x + \Delta x$.
Such a scheme should be sufficiently accurate, numerically stable, and should be able to deal with the drift induced by strong gradients in the diffusivity.
For the present application it is also important to minimize the number of calculations
of the spatial derivatives (or avoid them altogether) of the velocity field and the diffusivity.
Such derivatives tend to be noisy when they are determined from the raw output of a MHD code.
In the case of the calculation of the momentum change this can be achieved by using the method proposed by Marcowith \& Kirk (1999). In terms of $y = \ln(p/mc)$
one replaces the second equation of (\ref{CES}) by:
\be
\Delta y = - \frac{\Delta t}{3 \Delta x} \left[ V(x + \Delta x) - V(x) \: \right] \equiv \overline{\omega} \: \Delta t \; ,
\ee
assuming a steady flow for simplicity.
The position change $\Delta x$ is determined in the way outlined below.
This algorithm essentially uses the spatial average of the acceleration rate along the orbit of a simulated particle. For a steady flow:
\be
\label{omave}
\overline{\omega} = \frac{1}{\Delta x} \int_{x}^{x + \Delta x} {\rm d}x \: \left[ - \frac{1}{3} \frac{{\rm d} V(x)}{{\rm d} x} \: \right] =
- \frac{V(x + \Delta x) - V(x) }{3 \Delta x} \; .
\ee
For time-varying flows (where ${\rm d}V/{\rm d}x \: \Longrightarrow \: \partial V/\partial x$),
and in the unlikely case that the advective step and the diffusive step cancel each other (so that $\Delta x = 0$)
this prescription can lead to singular behaviour.
However, this is easily caught in a numerical scheme and correctly dealt with by putting $\Delta y = 0$ in that case.
\section{A predictor-corrector scheme for spatial transport}
We have found that a scheme
proposed by Kloeden \& Platen (1992) gives excellent results in the case of strong drift due to gradients in the diffusivity,
where the CES fails to be sufficiently accurate. It is a second-order predictor corrector scheme, called the KPPC scheme in what follows.
In particular this scheme improves the accuracy of the spatial transport of the particles, which leads to a better sampling of the
acceleration rate $\omega(x)$ in the shock compression.
For the problem at hand this scheme takes the following form:
\subsubsection*{Step 1: first supporting position value}
As a first step one calculates a first supporting position value $\tilde{x}$ using the Cauchy-Euler scheme:
\be
\label{s1step}
\tilde{x} = x + U(x \: , \: t) \: \Delta t + \sqrt{2 D(x) \Delta t} \: \xi_{t}
\ee
For simplicity we adopt a constant time step $\Delta t$. The stochastic variable
$\xi_{t} \in {\rm N}(0 , 1)$ is drawn from a normal distribution with zero mean and unit dispersion.
Two additional supporting position values are calculated,
\be
\label{s2step}
x_{\pm} = x + U(x \: , \: t) \: \Delta t \pm \sqrt{2 D(x) \Delta t} \; ,
\ee
that correspond with the position of two hypothetical particles that experience $\pm$ the rms diffusive step.
The stochastic variable $\xi_{t}$ is now {\em fixed} at the value used in (\ref{s1step}). It does not
change in any of the subsequent steps of the algorithm that are outlined below.
\subsubsection*{Step 2: predicted position value}
As a next step one calculates the predicted position $\bar{x}$ at time $t + \Delta t$ as
\be
\label{pstep}
\bar{x} = x + \bar{U} \: \Delta t + \overline{\Delta x}_{\rm diff}(\xi_{t}) \; .
\ee
The mean velocity $\bar{U}$ used in the advective + drift term is an average velocity, defined by using the first supporting value $\tilde{x}$:
\be
\label{barU}
\bar{U} = \half \left[ U(x) + U(\tilde{x}) \: \right] \; .
\ee
The improved stochastic diffusive step $\overline{\Delta x}_{\rm diff}(\xi_{t})$ equals
\ba
\label{barDiff}
\overline{\Delta x}_{\rm diff}( \xi_{t}) & = & \left( \frac{\Delta x_{\rm diff}^{+} + \Delta x_{\rm diff}^{-} + 2 \Delta x_{\rm diff}}{4} \right) \: \xi_{t} \nonumber \\
& & \\
& + & \left( \frac{\Delta x_{\rm diff}^{+} - \Delta x_{\rm diff}^{-}}{4} \right) \: \left( \xi_{t}^2 - 1 \right) \; . \nonumber
\ea
Here $\Delta x_{\rm diff}$ is the rms diffusive step (\ref{Dstep}) at the old position and $\Delta x_{\rm diff}^{\pm}$ corresponds to the rms diffusive step evaluated
at the two supporting positions $x_{\pm}$ that were defined in relation (\ref{s2step}):
\be
\Delta x_{\rm diff}^{\pm} \equiv \sqrt{2 D(x_{\pm}) \Delta t} \; .
\ee
The second term of the corrected diffusive step (\ref{barDiff}) corrects for the
`lopsidedness' of the random walk that results from the gradient in the diffusivity.
\subsubsection*{Step 3: corrector step and final position}
One finally obtains the corrected (and final) position $x_{\rm c} \equiv x(t + \Delta t)$ at time $t + \Delta t$ in the following way:
\be
\label{cstep}
x_{\rm c} = x + \half \left[ U(x) + U(\bar{x}) \: \right] \: \Delta t + \overline{\Delta x}_{\rm diff}(\xi_{t}) \; .
\ee
This last step uses the same diffusive step as in the predictor cycle but corrects the advective step using the predicted position $\bar{x}$ (see Eqn. \ref{pstep}).
This is a reasonable approach as (on average) $\Delta x_{\rm diff} \gg \Delta x_{\rm adv}$.
It was already noted by Marcowith \& Kirk (1999) that a careful
treatment of the advective step (including drift) is more important for the accuracy of the scheme than the treatment of the diffusive step.
This scheme and the tests presented below bear that out.
A few remarks about the implementation of this scheme are in order.
First of all, this scheme is computationally about 6 times more expensive than the Cauchy-Euler scheme.
An order-of-magnitude increase of the computational effort is typical when switching from an explicit, first-order scheme to a second-order accurate
predictor-corrector
scheme or the closely related Runge-Kutta type schemes.
Secondly: for the term that corrects for diffusivity gradients to be effective one should {\bf not} employ an often-used
numerical approximation for the Wiener process that replaces the normal distribution for $\xi_{t}$ by a two-point distribution of values, choosing $\xi_{t} = \pm 1$,
where the two possible signs are drawn randomly with
equal probability ${\cal P}_{+} = {\cal P}_{-} = \half$. In that numerical approximation for the Wiener process
the second term in the expression (\ref{barDiff}) vanishes identically, and much of the scheme's
improved accuracy with respect to the Cauchy-Euler scheme is lost.
In that respect one might expect that a symmetric three-value scheme for $\xi_{t}$, for example
(in the notation $[value \: | \: probability]$)
\be
\label{threepoint1}
\xi_{t} \in \left[ - \sqrt{3} \: | \: \frac{1}{6} \: \right] \; \; , \; \; \left[ 0 \: | \: \frac{2}{3} \: \right] \; \; , \; \; \left[ + \sqrt{3} \: | \: \frac{1}{6} \: \right] \; ,
\ee
works better.
\section{Tests of the algorithm}
\subsection{Basic assumptions}
We have tested the KPPC scheme as described here, comparing its performance to the performance of the simpler Cauchy-Euler scheme.
For this test we use scaled (dimensionless)
variables where the fluid velocity $V$ is measured in units of the shock speed and position $x$ along the shock normal is in units of the shock thickness $L_{\rm s}$
For clarity we keep $L_{\rm s}$ in the equations even though $L_{\rm s} = 1$ in the numerical implementation.
The velocity $V(x)$ is in the direction of positive $x$, given by
\be
\label{Vprofile}
V(x) = \frac{r + 1}{2r} - \frac{r - 1}{2r} \: {\rm tanh}\left( \frac{x}{L_{\rm s}} \right) \; .
\ee
The velocity decreases with increasing $x$, from $V(-\infty) \equiv V_{1} = 1$ to $V(+ \infty) \equiv V_{2} = 1/r$. This means
that we work in the rest frame of the shock and measure the flow velocity in units of the shock velocity with respect of the upstream medium.
Here $r > 1$ is the compression ratio of the shock transition in the sense that (for this one-dimensional
steady flow) the conservation of mass implies $\rho V = {\rm constant}$, with $\rho$ the mass density. The density
contrast between the far upstream and far downstream state follows as
\be
\frac{\rho(\infty)}{\rho(- \infty)} \equiv \frac{\rho_{2}}{\rho_{1}} = \frac{V_{1}}{V_{2}} = r \; .
\ee
This velocity profile models the shock as a stationary and smooth transition, with a width (velocity gradient scale) $L_{\rm s}$.
To model a varying diffusion coefficient we adopt a diffusion coefficient that varies with position $x$ as
\be
\label{Dprofile}
D(x) = D_{1} \: \left[ \frac{\sigma + 1}{2 \sigma} - \frac{\sigma - 1}{2 \sigma} \: {\rm tanh} \left( \frac{x}{L_{\rm d}} \right) \: \right] \; .
\ee
Here $D_{1}$ is a constant dimensionless diffusivity that is related to the physical diffusivity $D_{\rm phys}$ far ahead of the shock
by $D_{1} = D_{\rm phys}/L_{s} V_{s}$ with
$V_{s}$ the shock velocity. The diffusivity decreases if one moves from upstream ($x < 0$) to downstream ($x > 0$)
across the shock, with a ratio of asymptotic values equal to
\be
\frac{D(- \infty)}{D(\infty)} \equiv \frac{D_{1}}{D_{2}} = \sigma \ge 1 \; ,
\ee
the kind of behavior one expects in astrophysical applications.
The scale length for the variation of the diffusion coefficient in these units is of order $L_{\rm d}$.
For future use we define the quantity
\be
\varepsilon \equiv \frac{L_{\rm s}}{L_{\rm diff}(- \infty)} = \frac{V_{1} L_{\rm s}}{D_{1}} \; .
\ee
This is essentially the {\em P\'eclet number} of the shock based on the cosmic ray diffusivity. In terms of this quantity one has
\be
\label{driftratio}
\left| \frac{\Delta x_{\rm drift}}{\Delta x_{\rm adv}} \right| \simeq \frac{\sigma - 1}{2 \sigma } \: \frac{L_{\rm s}}{\varepsilon L_{\rm d}} \; .
\ee
A sharp shock in the present context corresponds to $\varepsilon \ll 1$.
The main test of the algorithm lies in its ability to reproduce the spectrum predicted by the analytical theory of DSA at a
steady shock with $\varepsilon \ll 1$, $\sigma \simeq r$ and $L_{\rm s} \simeq L_{\rm d}$. In the limit of a infinitely thin shock with $\varepsilon =0$
(in physical terms: a shock thickness that is much smaller
than the scattering mean free path of the accelerating cosmic rays) and
in the absence of radiation losses the predicted shape of the spectrum is a power law in momentum, with an index $q$
that depends only on the compression ratio $r$ (e.g. Axford, Leer \& Skadron, 1977,
Bell, 1978; Blandford \& Ostriker, 1978). In present notation, using the momentum $p$ rather than $y = \ln(p/mc)$:
\be
\label{DSAlaw}
N(x \: , \: p) = \frac{{\rm d} {\cal N}}{{\rm d}x \: {\rm d} \ln p} \propto p^{-q} \; \; , \; \; q(\varepsilon = 0) = \frac{3}{r - 1} \; .
\ee
In our tests of the algorithm we have assumed a diffusion coefficient that is independent of particle momentum, so
this power-law behavior is valid uniformly across the grid, with only
the concentration of test particles varying with position $x$.
\subsection{Effect of finite shock thickness}
Since we assume a finite shock thickness, a situation typical of shocks obtained through numerical simulation, we need an expression for the slope
$q$ for finite $\varepsilon$. We use a perturbation analysis adapted from Drury (1983) and the closely related method of Schneider \& Kirk (1987).
The analysis presented below is valid when there is a small parameter $\varepsilon$, in this case the ratio of the shock thickness and the cosmic ray diffusion length:
\be
\label{epscond}
\varepsilon \equiv \frac{L_{\rm s}}{L_{\rm diff}} \ll 1 \; .
\ee
Consider the steady-state transport equation (\ref{Skeqn}) reformulated in terms of the
Vlasov distribution $f(\bm{x} \: , \: p) \equiv {\rm d}{\cal N}/({\rm d}^3 \bm{x} \: {\rm d}^3 \bm{p})$.
In our application we have $f(x \: , \: p) = N(x \: , \: p)/4 \pi p^3$.
Assuming a one-dimensional steady flow in the $x$-direction the equation for $f(x \: , \: p)$ with $\partial f/\partial t = 0$ reads:
\be
\label{Skeqn2}
V \: \frac{\partial f}{\partial x} - \frac{\partial}{\partial x} \left( D \frac{\partial f}{\partial x} \right) =
\frac{1}{3}\frac{{\rm d} V}{{\rm d} x} \: \left( p \frac{\partial f}{\partial p} \right) \; .
\ee
Schneider \& Kirk use a variant of the Ricatti transformation (e.g. Polyanis \& Manzhirov, 2007, Ch. 12.2), which we slightly modify here.
We introduce a dimensionless position variable $X$,
\be
X(x) \equiv \int_{0}^{x} \: \frac{V(x') {\rm d} x'}{D(x')} \; ,
\ee
and define (c.f. Schneider \& Kirk, 1987)
\be
G(x \: , \: p) = -\frac{\displaystyle D \: \frac{\partial f}{\partial x}}{f(x \: , \: p)} = -\frac{\displaystyle V \: \frac{\partial f}{\partial X}}{f(x \: , \: p)}\; .
\ee
Schneider \& Kirk also assume a power-law momentum dependence,
\be
\label{powerl}
f( x \: , \: p) \propto p^{-\qbar} \; ,
\ee
which can only be strictly justified if the diffusion coefficient is independent of momentum, the case we consider here, and for momenta well above the injection momentum.
Note that we have $\qbar = q + 3$.
In that case $G(x \: , \: p)$ is a function $G(X)$ of position alone.
Equation (\ref{Skeqn2}) can be written as a non-linear ordinary differential equation for $G(X)$:
\be
\label{Reqn}
\frac{{\rm d}G}{{\rm d}X} + \frac{\qbar}{3} \: \frac{{\rm d}V}{{\rm d}X} = G + \frac{G^2}{V} \; .
\ee
This is the relation derived by Schneider \& Kirk (1987), generalized to the case of a position-dependent diffusion coefficient.
The boundary conditions for $G(X)$ at $X = \pm \infty$ are
\be
\label{BCs}
G(- \infty) = - V(-\infty) \equiv - V_{1} \; \; , \; \; G(+ \infty) = 0 \; .
\ee
The first condition assumes that there are no pre-existing particles far ahead of the shock so that asymptotically $f(x \: , \: p) \propto {\rm exp}(X)$ \
for large negative $X$, where $V(X) \simeq V_{1}$ is approximately constant.
The second condition, which states that the diffusive contribution $\propto \partial f/\partial x$
to the flux vanishes asymptotically far behind the shock, ensures
that the particle density remains finite as $X \: \longrightarrow + \infty$.
We now assume that condition (\ref{epscond}) is satisfied, meaning that the shock transition must be sufficiently sharp.
In that case, the left-hand side of Eqn. (\ref{Reqn}) will be large in a region $\Delta X \sim \varepsilon$ when compared with the two terms on the right-hand side.
This behavior can be formalized by using $\varepsilon$ as a formal ordering parameter, replacing (\ref{Reqn}) by
\be
\label{Reqn2}
\frac{1}{\varepsilon} \left( \frac{{\rm d}G}{{\rm d}X} + \frac{\qbar}{3} \: \frac{{\rm d}V}{{\rm d}X} \right) = G + \frac{G^2}{V} \; .
\ee
We seek solutions of the form
\be
\label{Gexp}
G(X) = G_{0}(X) + \varepsilon \: G_{1}(X) + \varepsilon^2 \: G_{2}(X) + \cdots
\ee
and expand the slope as
\be
\label{qexp}
\qbar = \qbar_{0} + \varepsilon \: \qbar_{1} + \varepsilon^2 \: \qbar_{2} + \cdots
\ee
We can now solve (\ref{Reqn2}) at each order of $\varepsilon$, putting $\varepsilon = 1$ at the end of the calculation.
At leading order ($\varepsilon^{-1}$) one has
\be
\label{zeroG}
\frac{{\rm d} G_{0}}{{\rm d}X} + \frac{\qbar_{0}}{3} \: \frac{{\rm d}V}{{\rm d}X} = 0 \; ,
\ee
subject to the boundary conditions (\ref{BCs}) for $G_{0}(X)$:
\be
G_{0} (- \infty) = - V_{1} \; \; \mbox{and} \; \; G_{0}(+ \infty) = 0 \; .
\ee
The solution is elementary:
\be
\label{Gzero}
G_{0}(x) = \frac{\qbar_{0}}{3} \: \left[ \: V_{2} - V(X) \: \right] \; ,
\ee
with $V_{2} = V(+ \infty)$ the asymptotic velocity downstream.
The zero-order slope $\qbar_{0}$ can be found by integrating (\ref{zeroG}) from $X = - \infty$ to $X = + \infty$ and using the boundary conditions.
Another elementary calculation gives
\be
\frac{\qbar_{0}}{3} \: \left(V_{2} - V_{1} \right) + V_{1} = 0 \; .
\ee
One finds (as expected) that $\qbar_{0}$ is the slope associated with an infinitely thin shock, where the velocity jumps from $V_{1}$ to $V_{2} < V_{1}$ at $x = 0$:
\be
\label{qzero}
\qbar_{0} = \qbar(\varepsilon = 0) = \frac{3 V_{1}}{V_{1} - V_{2}} \; .
\ee
This also ensures the the boundary condition at $X = - \infty$ is satisfied as (\ref{qzero}) substituted into (\ref{Gzero})
implies
\be
\label{G0sol}
G_{0}(X) = - \frac{V_{1} \: \left( V(X) - V_{2} \right)}{V_{1} - V_{2}} \; .
\ee
At next order ($\varepsilon^{0}$) one has:
\be
\label{firstG}
\frac{{\rm d} G_{1}}{{\rm d}X} + \frac{\qbar_{1}}{3} \: \frac{{\rm d}V}{{\rm d}X} = G_{0} + \frac{G_{0}^2}{V} \; ,
\ee
subject to the boundary condition
\be
\label{BC1}
G_{1}(- \infty) = G_{1}(+ \infty) = 0 \; .
\ee
Integrating (\ref{firstG}) from $X = - \infty$ to $X = +\infty$ immediately yields a relation for $\qbar_{1}$:
\be
\frac{\qbar_{1}}{3} \: \left( V_{2} - V_{1} \right) = \int_{- \infty}^{+ \infty} {\rm d}X \: \left(G_{0} + \frac{G_{0}^2}{V} \right) \; .
\ee
Solving for $\qbar_{1}$, using (\ref{qzero}) and (\ref{G0sol}):
\be
\label{qone}
\qbar_{1} = \qbar_{0} \: \int_{- \infty}^{+ \infty} {\rm d}X \:
\frac{\displaystyle V_{2} \left(V_{1} - V \right) \left(V - V_{2} \right)}
{\displaystyle V \: \left(V_{1} - V_{2} \right)^2} \; .
\ee
The function $G_{1}(X)$ is
\ba
\label{G1}
G_{1}(X) & = & \frac{\qbar_{1}}{3} \: \left( V_{1} - V \right) \nonumber \\
& & \\
& & \; \; \; \; \; - \int_{- \infty}^{X} {\rm d}X' \:
\frac{\displaystyle V_{1} V_{2} \left(V_{1} - V' \right) \left(V' - V_{2} \right)}
{\displaystyle V' \: \left(V_{1} - V_{2} \right)^2} \; . \nonumber
\ea
Here $V' \equiv V(X')$. Note that $\qbar_{1}$ and $G_{1}(X)$ vanish automatically if one uses the step function velocity profile of an infinitely thin shock,
with $V(X) = V_{1}$ for $X < 0$ and $V(X) = V_{2}$ for $X \ge 0$.
\nskip
At order $\varepsilon$ one finds the following equation for $G_{2}(X)$:
\be
\label{secondG}
\frac{{\rm d} G_{2}}{{\rm d}X} + \frac{\qbar_{2}}{3} \: \frac{{\rm d}V}{{\rm d}X} = G_{1} \: \left(1 + \frac{2 G_{0}}{V} \right) \; ,
\ee
subject to the boundary condition $G_{2}( - \infty) = G_{2}(+ \infty) = 0$. Integrating (\ref{secondG}) from $X = - \infty$ to $X = +\infty$
yields an equation for $\qbar_{2}$:
\ba
\label{qtwo}
\qbar_{2} & = & - \frac{3}{V_{1} - V_{2}} \int_{- \infty}^{+ \infty} {\rm d}X \: G_{1}(X) \: \left(1 + \frac{2 G_{0}(X)}{V(X)} \right)
\nonumber \\
& & \\
& = &
\frac{3}{V_{1} - V_{2}} \int_{- \infty}^{+ \infty} {\rm d}X \: G_{1}(X) \:
\left(
\frac{\left( V_{1} + V_{2} \right)V(X) - 2V_{1} V_{2}}{V (X)\left( V_{1} - V_{2} \right)} \right)\; .
\nonumber
\ea
This procedure can be extended to higher order, but little is gained at the expense of increasingly complex mathematics.
\subsubsection{Specific examples}
We use two examples of immediate importance for a test of the numerical scheme advocated here. Table 1 gives the parameters as used in the
numerical simulations presented below.
\subsubsection*{Model 1: hyperbolic tangent velocity profile and constant diffusivity}
The first example is the case of a uniform diffusivity $D(x) = D_{1}$, which formally corresponds to $\sigma = 1$ and $L_{\rm d} = \infty$.
This is the case where the CES is known to yield good results. This case has been treated before by Axford, Drury \& Summers (1982), who show that for
the hyperbolic tangent velocity profile (\ref{Vprofile}) adopted here the cosmic ray transport equation can be solved analytically.
They find an asymptotic slope of the momentum distribution equal to
\be
\label{DASq}
\qbar = \qbar_{0} \: \left( 1 + \frac{V_{2} L_{\rm s}}{2 D_{1}} \right) \; .
\ee
Here $\qbar_{0} = 3r/(r - 1)$.
The perturbation expansion used here (and in a slightly different form by Drury (1983)) reproduces this result. The hyperbolic tangent velocity law (\ref{Vprofile})
implies
\be
\label{Vder}
\frac{{\rm d}V}{{\rm d}x} = - \frac{2}{L_{\rm s}} \: \frac{\left( V_{1} - V \right) \left( V - V_{2} \right)}{V_{1} - V_{2}} \; .
\ee
Substituting this into the generally valid expression (\ref{qone}), together with ${\rm d}X = V \: {\rm d}x/D_{1}$, one finds:
\be
\qbar_{1} = - \frac{\qbar_{0} V_{2} L_{\rm s} }{2 D_{1} (V_{1} - V_{2}) } \: \int_{-\infty}^{\infty} {\rm d} x \: \left( \frac{{\rm d}V}{{\rm d} x} \right) =
\qbar_{0} \: \frac{V_{2} L_{\rm s}}{2 D_{1}} \; .
\ee
This agrees with result (\ref{DASq}) of Drury et al (1982). Using this in relation (\ref{G1}) one finds that $G_{1}(X) = 0$, which implies
$G_{n} = q_{n} = 0$ for $n \ge 2$. Here the perturbation expansion breaks off at order $\varepsilon$ and yields the {\em exact} asymptotic
result, as noted before by Drury (1983). In terms of the compression ratio $r = V_{1}/V_{2}$, the diffusion length far upstream $L_{\rm diff}^{-\infty} = D_{1}/V_{1}$ and
$q = \qbar - 3$ one has:
\be
\label{qDAS}
q = \frac{3}{r - 1} \: \left( 1 + \frac{V_{1} L_{\rm s}}{2 D_{1}} \right) = \frac{3}{r - 1} \: \left(1 + \frac{\varepsilon}{2} \right) \; .
\ee
\subsubsection*{Model 2: hyperbolic tangent profile and constant diffusion length}
As a second example we consider the case of a constant diffusion length:
\be
L_{\rm diff} = \frac{D(x)}{V(x)} = \frac{D_{1}}{V_{1}}\; .
\ee
This example is important as a test case of the predictor-corrector algorithm used here as it has a strong gradient in the diffusivity.
Formally it corresponds to $\sigma = r$ and $L_{\rm d} = L_{\rm s}$ so that in the shock
\be
\left| \frac{\Delta x_{\rm drift}}{\Delta x_{\rm adv}} \right| \simeq \frac{r - 1}{2 r \varepsilon} \; ,
\ee
which becomes large if $\varepsilon \ll 1$ for thin shocks.
If one adopts the hyperbolic tangent profile (\ref{Vprofile})/(\ref{Vder}) and uses the fact that
\be
{\rm d}X = \frac{{\rm d}x}{L_{\rm diff}} \; ,
\ee
relation (\ref{qone}) yields:
\ba
\qbar_{1} & = & - \: \frac{\qbar_{0} L_{\rm s}}{2 L_{\rm diff}} \: \left( \frac{V_{2}}{V_{1} - V_{2}} \right) \int_{- \infty}^{\infty} {\rm d} x \:
\left( \frac{1}{V} \frac{{\rm d}V}{{\rm d}x} \right) \nonumber \\
& & \\
& = &
\frac{\qbar_{0} L_{\rm s}}{2 L_{\rm diff}} \: \left( \frac{V_{2}}{V_{1} - V_{2}} \right) \: \ln \left( \frac{V_{1}}{V_{2}} \right) \; . \nonumber
\ea
The function $G_{1}$ can be calculated, but the integral over $G_{1}$ that determines the next order correction $\qbar_{2}$ to the slope can not be expressed in
elementary functions. Limiting ourselves to the first-order correction one has in terms of $q = \qbar - 3$ and $r = V_{1}/V_{2}$:
\be
q \simeq \frac{3}{r - 1} \: \left(1 + \frac{ r \: (\ln r) \: L_{\rm s}}{2(r - 1) L_{\rm diff}} \right) =
\frac{3}{r - 1} \: \left(1 + \frac{ r \: (\ln r) \: \varepsilon}{2(r - 1)} \right) \; .
\ee
Here the steepening due to a finite shock width $\sim L_{\rm s}$ is more pronounced than in the previous cases as the diffusion length near the shock transition is smaller
compared to the first case.
For instance: in a shock with compression ratio $r= 4$, the value expected for a strong shock in a mono-atomic gas, one has
$r \: \ln r/2(r - 1) \simeq 0.924$. We have obtained the second-order term for the important case $r = 4$ through numerical integration
of the integral in the expression for $\qbar_{2}$. We find:
\ba
\label{ris4}
q(r = 4) & = & 1 + 0.924 \: \frac{L_{\rm s}}{L_{\rm diff}} + 0.095 \: \frac{L_{\rm s}^2}{L_{\rm diff}^2} \nonumber \\
& & \\
& = &
1 + 0.924 \: \varepsilon + 0.095 \: \varepsilon^2 \; . \nonumber
\ea
Note that the end result in both cases is a series in $\varepsilon = L_{\rm s}/L_{\rm diff}$.
It should be pointed out that the term $\propto \varepsilon^2$ is small in the
second case, and vanishes completely in the first case. The smallness of the second-order correction to $q$
seems to be a rather general property of this expansion in $\varepsilon$ for reasonable velocity profiles.
As a further example: in the case of a linear velocity profile, where
\be
\label{linprofile}
V(x) = \left\{ \begin{array}{ll}
V_{1} & \mbox{for $x < - L_{\rm s}/2$,} \\
& \\
{\displaystyle \frac{V_{1} + V_{2}}{2} - \frac{V_{1} - V_{2}}{L_{\rm s}}} \: x & \mbox{for $-L_{\rm s}/2 \le x \le L_{\rm s}/2$,} \\
& \\
V_{2} & \mbox{for $x > L_{\rm s}/2$,} \\
\end{array}
\right.
\ee
and for a constant diffusion coefficient $D = D_{1}$ ($\sigma = 1$, $L_{\rm d} = \infty$), the same procedure yields:
\be
q = \qbar - 3 \simeq \frac{3}{r - 1} \left( 1 + \frac{\varepsilon}{6} + \frac{(r +1) \varepsilon^2}{360 r} \right) \; .
\ee
For $r = 4$ this is
\be
q(r = 4) = 1 + \frac{\varepsilon}{6} + \frac{\varepsilon^2}{288} \; .
\ee
These three examples suggest that the results obtained here for the slope $q(\varepsilon)$ are applicable even for $\varepsilon \simeq 1$.
The simulations presented in the next Section bear this out.
\begin{table}
\caption{Model parameters}
\label{table1}
\centering
\begin{tabular}{c c c c c} \hline \hline
& & & & \\
{\bf Model} & $r$ & $\sigma$ & $L_{\rm d}/L_{\rm s}$ & $V_{1} \: \Delta t/L_{\rm s}$\\
& & & & \\ \hline
& & & & \\
1 & 4 & 1 & $\infty$ & 0.05\\
& & & & \\
2 & 4 & 4 & 1 & 0.05\\
& & & & \\ \hline \hline
\end{tabular}
\end{table}
\subsection{Numerical results}
The two figures below show the results of numerical simulations for the two cases listed in Table 1. The approach is similar to the one employed by
Achterberg and Kr\"ulls (1992): particles are injected close to the shock, and followed over a grid that extends from $X = - 10$ to $X = + 10$.
Particles are detected as they cross the downstream boundary $X_{\rm max} \simeq V_{2} x_{\rm max}/D_{2} = 10$, which acts as an absorber.
The influence of a downstream absorbing boundary
on the slope $q$ decays as ${\rm exp}(-X_{\rm max})$ with respect to unity, and is negligibly small for these parameters.
Particle splitting is used at intervals equidistant in
$\log p$ to minimize the effects of Poisson noise at large momenta where fewer particles reside in the distribution.
The spectra obtained in this way are strict power laws that extend over five
decades ($\Delta y \simeq 11.5$) in particle momentum. We use a fixed time step that corresponds to $V_{1} \: \Delta t = 0.05 \: L_{\rm s}$,
so that the advective step resolves the shock transition.
In practice, good results are obtained if $\Delta x_{\rm adv} \lse 0.1 \: L_{\rm s}$.
The diffusion coefficient
varies from $D_{1} = 1$ to $D_{1} = 100$, which corresponds to a diffusive step in the range $\Delta x_{\rm diff} \simeq 0.3 - 3$. Note that
in our implementation we have scaled the spatial coordinate $x$ with the shock width so that $L_{\rm s} = 1$.
\begin{figure}
\includegraphics[width=84mm]{figure1.eps}
\caption[]{Results for Model 1: the case of a hyperbolic tangent velocity profile and constant diffusivity.
The solid curve gives the result (\ref{qDAS}), the open stars are the results of the KPPC scheme and the solid stars
the results obtained using the CES. The shock compression ratio equals $r = 4$, the case of a strong hydrodynamical shock.
The parameter $\varepsilon = L_{\rm s}/L_{\rm diff}$ varies from 0.01 to unity.
Note that the figure employs a logarithmic scale for $\varepsilon$.}
\label{fig1}
\end{figure}
\begin{figure}
\includegraphics[width=84mm]{figure2.eps}
\caption[]{Results for Model 2: the case of a hyperbolic tangent velocity profile and constant diffusion length $L_{\rm diff} = D(x)/V(x)$.
The solid curve gives the result (\ref{ris4}), the open stars are the results of the KPPC scheme and the solid stars
the results obtained using the CES. The shock compression ratio equals $r = 4$, the case of a strong hydrodynamical shock.
Again a log-linear scale is employed in this figure. Note that the vertical scale differs from the one used in Figure 1.}
\label{fig2}
\end{figure}
Figure 1 shows the results obtained for the case of a constant diffusion coefficient (Model 1). In this model there is no drift term.
In this case the CES and the KPPC scheme give comparable results that closely follow the theoretical prediction up to $\varepsilon = L_{\rm s}/L_{\rm diff} = 0.04$.
For smaller values of $\varepsilon$ (i.e. larger values of $D_{1}$) the both schemes become inaccurate as they under-sample the acceleration rate $\omega(x)$ in the shock transition.
Figure 2 shows the results for Model 2, where $L_{\rm diff} = D(x)/V(x)$ is kept constant so that $L_{\rm s} = L_{\rm d}$ and $\sigma = r = 4$.
This is the model with a large drift in the shock due to the
gradient in the diffusivity. The results for the slope of the momentum distribution obtained using the CES (the filled stars) deviate significantly from the slope
obtained using the perturbation expansion if $\varepsilon = L_{\rm s}/L_{\rm diff} = L_{\rm d}/L_{\rm diff} < 0.1$. In contrast, the KPPC scheme gives significantly
better results that are usable up to $\varepsilon \simeq 0.02$. The error in the value of the slope $q$ returned by the KPPC scheme is typically 3 times smaller than
the error produced by the CES.
The deviation from the analytical result becomes significant if the magnitude of the drift term in the statistically sharp spatial
step $\Delta x_{\rm s}$ becomes of the same order as the shock thickness:
\be
\left| \Delta x_{\rm drift} \right| = \left| \frac{{\rm d} D}{{\rm d}x} \right| \: \Delta t \ge L_{\rm s} \; .
\ee
Making the estimate $|{\rm d}D/{\rm d}x| \simeq |\Delta D|/L_{\rm s}$ as it applies to typical situations where the diffusion coefficient jumps by an amount
$\Delta D$ across the shock, the accuracy of the KPPC scheme is lost if
\be
\frac{\Delta x_{\rm drift}}{L_{\rm s}} = \frac{|\Delta D|}{D_{1}} \frac{V_{1} \Delta t}{\varepsilon L_{\rm s}} > 1 \; .
\ee
For instance: in the results shown in Figure 2 the deviation in the slope returned by the KPPC scheme becomes large when $|\Delta x_{\rm drift}| \simeq 2 L_{\rm s}$.
In our test of the KPPD scheme we have assumed that the diffusion coefficient is independent of momentum. In many astrophysical applications one expects the mean free
path to increase with momentum so that the diffusion coefficient scales as $D \propto p^{\alpha}$. For instance: one often assumes {\em Bohm Diffusion} with a mean-free-path
equal to the gyro radius, $\lambda_{\rm mfp} = r_{\rm g} \simeq pc/qB$ where $q$ is the particle charge. For relativistic particles ($v\sim c$) this implies $D \propto p$.
The KPPD scheme should be able to give reliable result in this case also, as long as the time steps are such that $|\Delta x_{\rm drift}| \lse L_{\rm s}$.
The scaling $D \propto p^{\alpha}$ implies $\Delta x_{\rm drift} \propto p^{\alpha}$, so it may be necessary, depending on the dynamic range in $p$,
to employ the scheme with smaller time steps for the high-energy particles
in the population than for the low-energy particles.
\section{Discussion and conclusions}
We have argued that the simulation of diffusive shock acceleration of cosmic rays through the numerical integration of stochastic differential equations
needs a more accurate, second-order scheme when large gradients in the cosmic ray diffusivity arise. This situation is astrophysically relevant for oblique shocks or for shocks
where the cosmic rays generate strong magnetic turbulence in the shock vicinity. The Kloeden-Platen predictor-corrector scheme proposed here
is such a scheme. We have demonstrated using simple simulations that the Kloeden-Platen predictor-corrector scheme is significantly more accurate for small shock widths
coupled with a strong gradient in the cosmic ray diffusivity.
For large shock widths (P\'eclet numbers larger than $\sim 0.25$), or when the gradient in the cosmic ray diffusivity is small, the two schemes produce comparable results
in terms of the accuracy of the momentum distribution that is obtained for the simulated particles.
Given the fact that the Kloeden-Platen predictor-corrector scheme is computationally about six times more expensive than the Cauchy-Euler scheme, one might consider
implementing a hybrid approach where one switches between the simpler Cauchy-Euler scheme and the Kloeden-Platen predictor-corrector scheme with the switch
based on the value of $|\Delta x_{\rm drift}/L_{\rm s}|$, the magnitude ratio of the drift term and the shock thickness
in the stochastic differential equation (\ref{CES}) and the shock width. The results obtained here suggest that the KPPC scheme
is needed for sufficient accuracy whenever $|\Delta x_{\rm drift}| \ga 4 \Delta x_{\rm adv}$ close to the shock.
We have checked whether results with similar accuracy can be achieved with the Cauchy-Euler scheme, simply by reducing the time step until
the computational expense is similar to that of the KPPC scheme with the larger time step.
This turns out not to be the case. As an example we consider Model 2 for the case $\varepsilon = 0.04$. The analytical estimate for the slope is
$q_{\rm th} = 1.037$. We ran both schemes with $V_{1} \Delta t = 0.1$, $0.05$, $0.025$ and $0.0125$ in units where $L_{\rm s} = 1$,
thus halving the time step each time. The KPPC scheme consistently returns (within errors due to Poisson noise)
a slope $q_{\rm KPPC} = 1.035$, quite close to the (approximate) theoretical result.
Table 2 gives the corresponding result for the slope obtained with the CES, $q_{\rm CES}$. The slope still has a sizable error
even for the smallest time step, 8 times smaller than the largest step where the KPPC scheme already performs satisfactorily.
The slope $q_{\rm CES}$, although decreasing as the time step gets smaller, remains consistently too large.
We conclude that, for given computational expense, the KPPC scheme is still superior.
\begin{table}
\caption{Performance Cauchy-Euler scheme}
\label{table2}
\centering
\begin{tabular}{c c c c c} \hline \hline
& & & & \\
Advective step $V_{1} \: \Delta t/L_{\rm s}$ & 0.1 & 0.05 & 0.025& $ 0.0125$\\
& & & & \\
Slope $q_{\rm CES}$ & 1.222 & 1.150 & 1.098 & 1.077 \\
& & & & \\ \hline \hline
\end{tabular}
\medskip
The slope $q$ of the simulated momentum distribution returned by the Cauchy-Euler scheme for Model 2 with $\varepsilon = 0.04$ for different time steps.\\
The theoretical slope equals $q_{\rm th} = 1.037$.
\end{table}
We note in passing that the Kloeden-Platen predictor-corrector scheme may also be useful
in other numerical applications of stochastic differential equations, such as the solution of the equations that describe stochastic particle
acceleration by waves (Fermi-II acceleration). Here the equation for the evolution of the momentum distribution of the accelerating
particles contains a momentum diffusion term that can be written in the form
(e.g. Melrose, 1980)
\be
\left( \frac{\partial f}{\partial t} \right)_{\rm acc} = \frac{1}{p^2} \frac{\partial}{\partial p} \left(p^2 D_{p} \: \frac{\partial f}{\partial p} \right) \; ,
\ee
and a large drift term is unavoidable. The relevant drift velocity (in this case corresponding to the mean momentum gain) is
\be
\left( \frac{{\rm d} p}{{\rm d}t} \right)_{\rm drift} = \frac{1}{p^{2}} \frac{\partial}{\partial p} \left(p^2 \: D_{\rm p} \right) \; .
\ee
In these expressions $D_{p}$ is the momentum diffusion coefficient.
|
\section{Introduction}
The control afforded by Feshbach
resonance phenomena in ultracold atomic gases has enabled the
exploration of strongly correlated degenerate Fermi systems. In
a recent study of the possibility of itinerant ferromagnetism
\cite{bloc,ston,zong,duin},
Jo \textit{et al.} \cite{jo} attempted to observe the physics
behind the Stoner model in an atomic gas of $^{6}$Li atoms.
Evidence for ferromagnetic ordering was seen.
This experiment has generated a great deal of theoretical
research \cite{lebl, cond, ChangHub2010, troy, triv, reca}. The results have been
debated as to whether a ferromagnetic transition or a strong
correlation effect was seen. Quantitative comparison with
experiment has not been achieved.
A key issue is to find an appropriate model for the experiment.
In the experiment a strong attractive interaction is quickly switched
on. Predictions of the critical ratio of interaction strength to
interatomic spacing for the ferromagnetic transition from mean field theory \cite{ston,lebl},
second order corrections \cite{duin, reca} and QMC calculations
\cite{cond, troy, triv} are on the order of $k_{F}a\sim 1$;
about two times lower than that from the Jo \textit{et al.}
experiment. In almost all calculations,
a positive interaction \cite{cond, troy} or a Jastrow factor with two-body nodes \cite{triv} is assumed,
using the scattering length approximation(SLA). Moreover, the various approaches differ in details
of the nature of the transition \cite{ston,duin}.
The two-body SLA neglects
the low-lying molecular states.
The zero-energy $s$-wave scattering length $a$ is defined by
the long distance form of the out-going scattering wave
\begin{equation}
\psi \left(r\rightarrow \infty\right) \propto \frac{\sin\left[
k(r-a) \right]}{kr} \stackrel{k\rightarrow
0}{\longrightarrow} 1 -\frac{a}{r}.
\end{equation}
For a contact (zero range) potential, $a$ is the radius of the
first wavefunction node: $\psi\left(r=a\right)=0$. The
SLA replaces the underlying atomic interaction
by a purely repulsive potential which has the same two-body
scattering length.
This is analogous to the idea of
pseudo-potentials in electronic structure. A pseudo-potential
can be generated in an atomic calculation to replace the strong
Coulomb potential of the nucleus and the effects of the tightly
bound core electrons by an effective ionic potential acting on
the valence electrons and then used to compute properties of
valence electrons in molecules or solids, since the core states
remain almost unchanged.
The approach is widely used in electronic structure calculations.
However, it leads to an inaccuate model
if a pseudopotential is used for systems compressed
to high density and the electron cores start to overlap.
Many experiments in cold
atomic systems are performed near Feshbach resonance where the
scattering length is comparable to
interatomic separation. In this situation, the lower-lying
molecular bound states giving rise to resonance can overlap,
causing the scattering
states to distort in order to remain orthogonal to the bound
states. With more experimental effort expected in the study of related systems,
precise and reliable comparisons from quantum simulations will be important.
Yet accurate many-body calculations will not be possible without a quantitative
understanding of the effective interactions and their effect on the different states.
Even the identification of the scattering state in a dense system requires explanation.
In this article, we quantify this effect by explicitly
including the molecular bound states and treating the interaction exactly.
We consider systems of
two, three and four spin-$\frac{1}{2}$ fermions interacting
through a contact interaction. The energy spectrum as a
function
of the two-body interaction strength
is obtained by using an exact
numerical method on a lattice and then extrapolated to the
continuum limit.
We show how the upper branch can be identified for a many-body system.
The properties of the nodal surface of the scattering many-body states are investigated.
To compare with the exact solutions, calculations are also made with the SLA,
by replacing the attractive contact interaction with a zero boundary
condition.
In both cases an adiabatic ferromagnetic transition is
stabilized. The SLA breaks down for large $a$, leading to
a severe underestimation (by almost a factor of 2) of the
transition point.
\section{Method}
\label{sec:method}
We consider a system of two-component
fermions moving in a periodic box with length $L$ to model a
gas of $^{6}$Li atoms with two hyperfine species at non-zero density. All lengths
are expressed in units of $L$ and all energies in units of
$K_{0}=\frac{\hbar^{2}}{2m}(\frac{2\pi}{L})^{2}$. In the case
$a\gg r_{0}$ (where $r_{0}$ is a measure of the effective interaction range),
the interatomic potential can be modeled as a regularized $\delta$-function:
\begin{equation}
V(\mathbf{r},\mathbf{r}')=\frac{4\pi\hbar^{2}a}{m}\delta(\mathbf{r}-\mathbf{r}')\frac{\partial}
{\partial|\mathbf{r}-\mathbf{r}'|}|\mathbf{r}-\mathbf{r}'|
\label{diracdelta}
\end{equation}
where $a$ is the zero-energy scattering length and $m$ is the
mass.
We solve the Schr\"{o}dinger equation by putting the system on
a lattice with $n$ points in each direction and recover the
continuum limit by extrapolation. We then
approximate the kinetic energy by two different discrete
Laplacian operators \cite{laplacian}: (1) the Hubbard model with nearest neighbor
hopping and (2) a long range hopping model including up to the next nearest
neighbors. We model the bare two-particle interaction by a point contact potential on the grid
\begin{equation}
V^{\footnotesize{\mbox{grid}}}(\mathbf{r},\mathbf{r}') = -\frac{U}{\Delta^{3}}\delta_{\mathbf{r},\mathbf{r}'},
\label{kronecker}
\end{equation}
where $\Delta = L/n$ is the grid spacing. Here $U>0$ is the
strength of the attractive interaction; on the repulsive side of resonance
$U>U_{\infty}$, we have positive scattering length for unpaired
atoms and the mapping relation between the grid and continuum
is \cite{cast}
\begin{equation}
\frac{m}{4\pi\hbar^{2}a} = \frac{1}{U_{\infty}}-\frac{1}{U},
\label{mapping}
\end{equation}
where the unitarity point $a\rightarrow\infty$ occurs at
$U^{-1}_{\infty}= (2\pi)^{-3}\int
d^{3}\mathbf{k}(2\epsilon_{\mathbf{k}})^{-1} = \gamma m /(\hbar^{2}\Delta)$. Here $\epsilon_{\mathbf{k}}$
is the single particle dispersion relation and $\gamma$ is a numerical constant defined by the discrete Laplacian.
For choice (1) above, $\gamma\approx 0.2527$; for choice (2), $\gamma\approx0.2190$. When only nearest neighbor hopping is included, our Hamiltonian is the standard attractive Hubbard model,
but scaled by $1/\Delta^2$. Note our $U$ value scales as $\Delta$, while in the notation
of the Hubbard model, $U_{\infty}$ is a constant.
In the SLA, $U$ has the opposite sign. In particular,
when the scattering length $a$ is large, Eq.~(\ref{kronecker}) is replaced by a
hard-sphere potential with radius $a$.
If $a$ is smaller than the grid spacing, a finite but negative value of $U$ can be used in
the SLA, leading to the repulsive Hubbard model,
which clearly has a different strongly interacting or
large $a$ limit \cite{ChangHub2010} from that of Eq.~(\ref{kronecker}).
To determine the eigenvalues and eigenstates, we
start from a set of random trial states
$|\psi_{\alpha}^0\rangle$ where $1\le\alpha\le M$, and evolve the
states $|\psi^{i+1}\rangle = (\mathbf{1}-\tau
\hat{H})|\psi^{i}\rangle$. At each step of the evolution, the state vectors
are properly symmetrized and orthogonalized. As $i\rightarrow
\infty$, the states converge to the lowest $M$ eigenstates of
the Hamiltonian $\hat{H}$ within a given symmetry sector. The
errors are controlled and can be
reduced arbitrarily with increasing the number of grid points
or number of iterations. As discussed below, the computational
cost grows rapidly with system size, but significant reduction
can be achieved by invoking symmetries.
To assess the accuracy of iterative
diagonalization, we test the method on a two-particle problem.
The energy of the lowest two-particle scattering state is plotted in Fig~(\ref{fig:tbp}) as a
function of the dimensionless scattering length $k_{F}a$, where
$k_{F}=(3\pi^{2}\rho)^{1/3}$ is the Fermi wave vector and
$\rho$ the particle density. Both discrete representations of
the kinetic energy operator were used and they converge to the same
continuum limit: $n\rightarrow\infty$; the long-range hopping
is found to be less sensitive to the lattice spacing. Solving
the two-particle problem also enables
us to construct repulsive pseudo-potentials in the SLA by
inverting the 2-particle Schr\"{o}dinger equation.
\begin{figure}
\scalebox{0.3}[0.3]{\includegraphics[0,0][30cm,22cm]{Fig01.eps}
\caption{\label{fig:tbp} (Color online) The two-body $s$-wave
scattering energy $E_{1}$ as a function of $k_{F}a$
for grid sizes: $n=6,8,10,12,20,40$. The inset shows the
scaling with respect to $1/n$ at $k_{F}a=2.0$ with grid sizes
up to $n=90$. The long-range hopping model converges to the
continuum limit faster than the Hubbard model.}
\end{figure}
\section{Two fixed point potentials}
\label{sec:1body}
The simplest case where the scattering length approximation may
fail is the scattering of a single particle off of two fixed
contact potentials. This problem can be solved exactly in
infinite space \cite{brue}. The nodal surface of the
zero-energy scattering state is given as the solution of:
\begin{equation}
\frac{1}{|\mathbf{r}-\mathbf{R}_{1}|}+\frac{1}{|\mathbf{r}-\mathbf{R}_{2}|}= \frac{1}{a} + \frac{1}{|\mathbf{R}_{1}-\mathbf{R}_{2}|},
\label{exactnodes}
\end{equation}
where $\mathbf{R}_{1}$ and $\mathbf{R}_{2}$ are the location of
the two fixed scatterers. In the SLA, one would model the state
by the ground state with nodes defined by
$|\mathbf{r}-\mathbf{R}_{1}|=a$ and
$|\mathbf{r}-\mathbf{R}_{2}|=a$. The nodal surfaces described by Eq.~(\ref{exactnodes})
are shown in Fig (\ref{fig:tnodes}) in comparison with corresponding spheres in SLA.
Clearly the deviation from SLA becomes
significant as $a\sim |\mathbf{R}_{1}-\mathbf{R}_{2}|$. In particular, the spherical surfaces
in SLA becomes infinitely large at unitarity limit while Eq.~(\ref{exactnodes}) gives
rise to a finite surface. This result suggests that introducing a node in the two-body Jastrow
factor in the form $f(r)\sim (1-a/r)$ is insuffcient to characterize the effective pairwise repulsion
on the upper-branch \cite{triv}, as further discussed below in Sec.\ref{sec:SLAvsExact}.
\begin{figure}
\scalebox{0.8}[0.8]{\includegraphics[0,0][9cm,7.8cm]{Fig02.eps}
\caption{\label{fig:tnodes} (Color online) Nodal surface for the scattering wavefunction in a potential
generated by two fixed particles located at $(\pm d/2,0,0)$ in infinite space with $a/d=1/10,1/3,1,5/2$ (expanding outward), where $d$ denotes the distance between the two fixed scatterers.
The solid (blue) lines correspond to the nodes in SLA and the dashed (red) lines to the exact nodes. SLA gives
a reasonable approximation to the nodal surface for $a/d<1$ but the deviation becomes significant for large scattering lengths.}
\end{figure}
To study the effect of the SLA at finite density, we solved the
same problem numerically in a finite periodic box. The results
are summarized in Fig~(\ref{fig:twocenter}). It can be seen
from the graph that at large scattering length (i.e. at high
density), the SLA significantly overestimates the scattering
energy for the 3-body problem, i.e. the effect of low-lying
molecular states cannot be ignored. The exact solution achieves
a lower energy by distorting the nodal surfaces away from the
union of two spheres required by the SLA. As we show below,
this also applies to a system of more fermions.
\section{Four-particle model}
Now consider a minimal model for the ferromagnetic transition:
four spin-$\frac{1}{2}$ atoms in a cube with periodic boundary
conditions and interacting with a contact potential.
The spin polarized state $\Psi(1234) = \psi_{A}(1234) \otimes
\left|\uparrow\uparrow\uparrow\uparrow\right\rangle$ has a
totally antisymmetric spatial part $\psi_{A}(1234)$: for a
contact interaction it is noninteracting with an energy of
$4K_{0}$ in a zero total momentum eigenstate that has
translational invariance.
On the other hand, there are two linearly independent spin
states with zero total spin $S^{2}=0$ corresponding to
unpolarized states:
\begin{eqnarray}
\nonumber \chi_{MS} \propto \left|\uparrow\uparrow\downarrow\downarrow\right\rangle
+\left|\downarrow\downarrow\uparrow\uparrow\right\rangle-\frac{1}{2}\big[\left|\uparrow\downarrow\right\rangle + \left|\downarrow\uparrow\right\rangle\big]\otimes\big[\left|\uparrow\downarrow\right\rangle + \left|\downarrow\uparrow\right\rangle\big],\\
\nonumber \chi_{MA} \propto \big[\left|\uparrow\downarrow\right\rangle - \left|\downarrow\uparrow\right\rangle\big]\otimes\big[\left|\uparrow\downarrow\right\rangle - \left|\downarrow\uparrow\right\rangle\big].\quad\quad\quad\quad\quad\quad\quad\quad\,\,\;
\end{eqnarray}
The wavefunction is a linear combination of the above two
states $\Psi(1234) =\nonumber \psi_{MA}\otimes \chi_{MS} + \psi_{MS}\otimes
\chi_{MA}$. The symmetries of $\psi_{MA}$ and $\psi_{MS}$ in
coordinate-space are determined by the total antisymmetry of
the complete wavefunction including spins and coordinates.
For a system of four particles on a grid with $n$ points in
each direction, the discretized configuration space has
$n^{12}$ grid points. Translational symmetries along the three
spatial directions reduce the size of the configuration space
by a factor of $n^{3}$. Cubic symmetry of the periodic box
reduces the number of independent wavefunction values by a
factor of $48$ and permutation symmetries give an additional
$24$-fold reduction.
We evolve pairs of
non-magnetic states $\{\psi_{MS},\psi_{MA}\}$ within the
reduced domain, and whenever the value of the wavefunctions on
a grid point outside the reduced domain is needed in
off-diagonal projections, the exterior point is mapped back
into the reduced domain by symmetry transformations.
\begin{figure}
\scalebox{0.29}[0.29]{\includegraphics[0,0][30cm,23cm]{Fig03.eps}
\caption{\label{fig:twocenter} (Color online) The scattering
energy of a particle moving in the potential generated by two
fixed particles located
at $x=L/4$ and
$x=3L/4$, with $y=L/2$ and $z=L/2$. The SLA is obtained by replacing each potential by a
hard sphere (zero boundary condition) with the same scattering
length. The SLA gives accurate energies in weakly-interacting
limit ($a/L<0.2$) but overestimates the scattering energy for
$a$ comparable with $L$.
The inset shows the nodal region $(|\psi_{i}|<10^{-4})$ for the scattering states with
$a/L=0.1\mbox{(black)},0.2\mbox{(red)},0.4\mbox{(blue)}$.
The surfaces become noticeably non-spherical for $a/L>0.1$. }
\end{figure}
The ferromagnetic transition is identified as the crossing
between the lowest singlet scattering state and the fully
ferromagnetic state. To investigate the effect of using the SLA
in multi-particle scattering process, the attractive contact
interaction is replaced by a zero boundary condition and the
resulting critical ferromagnetic density is estimated.
Fig~(\ref{fig:spectrum}) shows the energy spectrum of a
four-particle system for $n=10$ as a function of the coupling
coefficient $U$.
In this calculation, the lowest 30 states were followed. Note
that we only considered states with the same symmetries as the
ground state, i.e. with even parity with respect to reflection
in $x$, $y$ or $z$. The resulting energy levels can be
classified into three categories by their behavior at strong
coupling: two-molecule states, molecule-atom-atom states and
four-atom scattering states. Level avoiding \cite{land} can be
observed between states belonging to different categories.
\section{Identification of the scattering states}
\label{sec:Idscatter}
The formation of molecular bound states is characterized by the
binding energy diverging linearly as $U\rightarrow \infty$. In
particular, the ground state wavefunction can be approximately
written as $\psi_{0}(13)\psi_{0}(24) -\psi_{0}(14)\psi_{0}(23)$
where $\psi_{0}$ is the two-body bound state, and the ground
state energy is approximately twice the two-particle binding
energy. As seen in Fig. (3) the two-molecule states have an
energy slope ($\partial E / \partial U$) about twice as large
as the molecule-atom-atom states. As $U\rightarrow \infty$
molecules become tightly bound; their energy spacings can be
understood in terms of colliding molecules. For a lattice
model, in contrast to a continuum model, the greater the
internal binding energy, the greater the total mass of the
molecule \cite{matt}.
The scattering state of strongly repulsive atoms is an excited branch and all cold atom experiments performed in this regime are metastable. In Jo \textit{et al.} experiment, the magnetic field ramp ($\sim 4.5$ms) is much slower compared to the characteristic time
scale of atomic collisions $\hbar/k_{B}T_{F}\sim \mu$s, and marches toward the resonance from the repulsive side $a\gtrsim 0$. At low density or in the weakly interacting
regime, the four-atom scattering state approaches the noninteracting line $2K_{0}$ and the
SLA is an accurate approximation. But there are some difficulties in defining the
scattering state at high density or in the strongly interacting regime because of the level avoiding phenomena. As shown in the inset of Fig~(\ref{fig:spectrum}), if the coupling coefficient $U$ is
tuned toward the resonance $U_{\infty}$, it is energetically more favorable to jump through the
successive avoided crossings. Thus, the change in the scattering
energy due to an adiabatic tuning of the interaction can then be determined by following the excited branch curve. It is drawn in bold in Fig~(\ref{fig:spectrum}).
There is another way to identify the upper branch (scattering states) quantitatively by using the momentum distribution and the pairing order. First, consider the wavefunctions
for the relative motion of two particles interacting through a large scattering length
of Eq.~(\ref{diracdelta}).
The zero-energy scattering state in coordinate space $\psi(\mathbf{r})\propto r^{-1}-a^{-1}$
takes the form $\psi(\mathbf{k})\propto 4\pi k^{-2} - (2\pi)^{3}a^{-1}\delta(\mathbf{k})$
in momentum space and diverges at $\mathbf{k}=0$. By contrast, the bound state $\psi(\mathbf{r})\propto r^{-1}e^{-r/a}$
takes the form $\psi(\mathbf{k})\propto 4\pi (k^{2}+a^{-2})^{-1}$
in momentum space and remains finite at $\mathbf{k}=0$. The momentum distribution $n(\mathbf{k})$
for scattering states is different from bound states at $\mathbf{k}=0$: scattering states have a larger
fraction of particle occupation at $\mathbf{k}=0$.
We also consider the pairing order defined by:
\begin{equation}
g_{2}\equiv n \Big\langle
\sum_{i<j}\delta_{\mathbf{r}_{i},\mathbf{r}_{j}}
\Big\rangle_{\alpha}
\label{g2def}
\end{equation}
for each state $|\psi_\alpha\rangle$.
The quantity $g_2$ measures double occupancy, and is related to the energy slope:
\begin{equation}
\frac{\partial E_{\alpha}}{\partial U} = \Big\langle \frac{\partial
\hat{H}}{\partial U} \Big\rangle_{\alpha} =
-\frac{1}{\Delta^{2}L}g_{2}.
\label{dEdU}
\end{equation}
For the scattering state, double occupancy decreases monotonically as the interaction
strength is increased (see e.g.~Ref.~ \cite{ChangHub2010}).
Thus the scattering state in each lattice system is
characterized by a vanishing energy slope as $U\rightarrow \infty$,
\begin{equation}
g_2 \rightarrow 0, \quad \frac{\partial E_{\alpha}}{\partial U} \rightarrow 0,
\label{doublecount}
\end{equation}
as can be seen in Fig~(\ref{fig:spectrum}).
The pairing density is also related to the tail of the momentum distribution,
which describes the short range physics.
At large $k$, the momentum distribution takes the form $n(\mathbf{k}) \rightarrow C/k^{4}$,
where the coefficient $C$ is called the \textit{contact}
in the Tan relations \cite{tan}. In the continuum limit
$\Delta\ll a$, the contact $C$
can be related to the energy slope, and hence $g_2$,
through the adiabatic sweep theorem
$\frac{dE}{da^{-1}} = -\frac{\hbar^{2}}{4\pi m}C$, which in our system yields:
\begin{equation}
g_{2}= L\Big[\gamma - \frac{\Delta}{4\pi a}\Big]^{2}\,C,
\label{g2contact}
\end{equation}
where
$\gamma$ is the numerical constant appearing in the definition of $U_{\infty}$.
For bound states this gives a finite $g_2$ and a finite energy slope.
Thus, in addition to the momentum distribution at $\mathbf{k}=0$, we can identify the scattering state from
the other states by the magnitude of $g_{2}$: scattering states have
a smaller fraction of double occupation
$\mathbf{r}_{i}=\mathbf{r}_{j}$. Note that the contact $C$ measures
the local density of pairs \cite{braa}. The momentum distribution $n(\mathbf{k}=0)$ and the pair parameter $g_{2}$ are plotted in Fig~(\ref{fig:g2v}) as functions of the energy for $k_{F}a=0.8\sim 1.3$. Scattering states are, by definition, in the range $E/K_{0}>2$ and can be identified by the peaks of $n(\mathbf{k}=0)$ and low values of $g_{2}$.
\section{\label{sec:SLAvsExact}Comparison of the SLA and exact results}
The ferromagnetic transition
in the four-atom system occurs when the scattering state
energy equals the noninteracting energy, $4K_{0}$. For a $n=10$
grid, the transition occurs at $U/U_{\infty}\approx 1.07$.
\begin{figure}
\scalebox{0.3}[0.3]{\includegraphics[0,0][30cm,23.8cm]{Fig04.eps}
\caption{\label{fig:spectrum} (Color online) The energy
spectrum of the lowest 30 $s$-states of four fermions with
contact interactions on a 3D grid with $n=10$. The energy
levels can be classified into two-molecule states,
molecule-atom-atom states and four-atom states. The ferromagnetic transition can be identified as
the lowest scattering state (heavy dark line) crossing the horizontal
ferromagnetic line around $U/U_{\infty}\approx 1.07$. The inset shows an enlargement of the lowest scattering
state and the associated avoided crossings.}
\end{figure}
\begin{figure}
\scalebox{0.3}[0.3]{\includegraphics[0,0][30cm,23.6cm]{Fig05.eps}
\caption{\label{fig:g2v} (Color online) The momentum distribution $n(\mathbf{k}=0)$ and the pairing parameter $g_{2}$ of the lowest 30 $s$-states of four fermions with
contact interactions on a 3D grid with $n=10$ versus energy. Scattering states have, by definition, $E/K_{0}>2$ and can be identified by the large magnitude of $n(\mathbf{k}=0)$ and small magnitude of $g_{2}$ compared to the other states. The peak structure
at the scattering state diminishes as the interaction parameter $k_{F}a$ increases.}
\end{figure}
Shown in Fig~(\ref{fig:fourbody}) is the energy of the
four-particle unpolarized scattering state as a function of the
scattering length $k_{F}a$ on grids with $n=6,8,10,12$ and
their extrapolation to the continuum limit, $n=\infty$. Avoided
crossings between levels appear as kinks. The excited
scattering state from the solution of the four-particle
problem crosses the ferromagnetic line at $k_{F}a\approx 1.8$,
which is in remarkable agreement with the reported experimental value of $k_{F}a=1.9\pm 0.2$
\cite{jo}.
Also shown is the scattering energy using the SLA;
this gives a ferromagnetic transition at $k_{F}a\approx 1.08\sim 1.09$ for grid sizes $n=8,10,12$,
consistent with previous calculations \cite{ston, duin,
cond, troy, triv, reca}. The earliest Fixed-node diffusion Monte Carlo calculations employed
the repulsive P\"{o}schl-Teller potential ($k_{F}a\approx 0.86$) \cite{cond}, hard spheres or repulsive
soft spheres ($k_{F}a\approx 0.82$) \cite{troy} and included backflow effects ($k_{F}a\approx 0.96$) \cite{triv}. For attractive
interactions modeled by spherical square wells or attractive P\"{o}schl-Teller potential, either variational
Monte Carlo ($k_{F}a\approx 0.86$) \cite{troy} or fixed-node diffusion Monte Carlo \cite{triv} ($k_{F}a\approx 0.89$) calculates the upper-branch metastable state by
imposing a nodal condition in the many-body wave function. The nodal condition ensures
that the calculation consist of unbound
fermionic atoms and no dimers or other bound molecules,
by introducing a Jastrow factor in the form of the scattering solution
of the attractive potential corresponding to positive energy.
As shown in Sec.~\ref{sec:1body}, the nodal structure obtained this way deviates significantly
from the true nodes in the upper branch. This explains why
all these calculations gave results similar to those from repulsive potentials, and all of them reproduced
the predicted behavior of the mean field theory and second order corrections.
The discrepancy between
the critical values of $k_{F}a$ reflects the limitations of perturbation theory in the regime
of strong coupling. Compared to repulsive potentials, using Jastow factor with nodes and including backflow effects
for attractive potentials improves
the result by making nontrivial modifications to the nodal structure, but still gives answers not qualitatively different from the repulsive potential, and fails to reveal the inadequacy of the SLA.
\begin{figure}
\scalebox{0.3}[0.3]{\includegraphics[0,0][30cm,23.6cm]{Fig06.eps}
\caption{\label{fig:fourbody} (Color online) The four particle
scattering energy as a function of the scattering length. The
energy of the ferromagnetic state is shown as the dashed
horizontal line. The extrapolation to continuum is performed
based on calculations on grids with $n=6,8,10,12$, which exhibits
the $1^{\mbox{st}}$-order linear scaling $1/N$ to high accuracy. The
transition to the ferromagnetic phase occurs at $k_{F}a\approx
1.8$ while the SLA gives the transition at $k_{F}a\approx
1.08$. The inset shows the scaling with respect to $1/n$ near the transition point $k_{F}a=1.75$ and $1.8$.}
\end{figure}
These observations suggest that lower-lying molecular states are
responsible for delaying the formation of the ferromagnetic
phase. However, calculations with more than 4 atoms are needed
to determine finite size effects. Such calculations are not
feasible with the current method but might be possible with
stochastic methods.
\section{Dynamics of four-particle model}
Because of the limited lifetime of the strongly interacting
gas, however, the magnetic field ramp in experiment is not
adiabatic and can lead to different explanations\cite{pekk,zhai}.
A recent work \cite{cond2} takes into account
the effect of atom loss by including a fictitious three-body term in the
effective Hamiltonian of the Fermi gas and found that the critical
interaction strength required to stabilize the ferromagnetic state
increases significantly. A full $T$-matrix analysis \cite{pekk} suggests that the pairing
instability can prevail over the ferromagntic instability
and the experimentally measured atom loss rate can be qualitatively explained
in terms of the growth rate of the pairing order parameter after a quench.
Thus, it is an interesting problem to study the dynamics of the
pairing instability after a quench using the wave functions obtained in this
work. Since the contact $C$ is identified as the integral over space of the expectation value of a local
operator that measures the density of pairs \cite{braa}, we characterize the pairing instability by the
count of double occupancy $g_{2}$ in Eq (\ref{g2contact}).
To study the dynamics of the pair formation, we choose the initial state to be the
unpolarized four-particle ground-state in the noninteracting limit and expanded in the basis of the lowest $16$ eigenstates with the final interaction after the quench. The time evolution is then evaluated using the eigenstate expansion $|\psi(t)\rangle = \sum_{m}c_{m}e^{-\frac{i}{\hbar}E_{m}t}|\phi_{m}\rangle$.
The pairing density $g_{2}(t)=\langle \psi(t)|g_{2}|\psi(t) \rangle$ is used to characterize the pair formation and the atom loss into molecules. The evolution diagram of $g_{2}(t)$ is shown in Fig (\ref{fig:evolution}). The nonmonotonic dependence of the maximum value of $g_{2}(t)$ in the first oscillation period on the final scattering length $k_{F}a$ after the quench is in qualitative agreement with experiment and the full $T$-matrix theory \cite{pekk}.
\begin{figure}
\scalebox{0.3}[0.3]{\includegraphics[0,0][30cm,23.6cm]{Fig07.eps}
\caption{\label{fig:evolution} (Color online) The first oscillation period in the evolution of $g_{2}(t)$ after a quench from a non-interacting state to an interaction strength $k_{F}a$, where $\tau = \hbar/\epsilon_{F}$. The inset shows the nonmonotonic behavior of the maximum value of $g_{2}(t)$ in the first oscillation period as a function of the final interaction strength $k_{F}a$. This calculation in done for four particles on a grid with $n=12$.}
\end{figure}
\section{Summary}
In summary, we have assessed the accuracy of the scattering length approximation
at high density or strong interaction
$k_{F}a \gtrsim 1$. It is demonstrated that if molecular states
mix with excitations, non-magnetic states are stabilized.
Identification of the upper branch in many-body calculations is discussed.
The corresponding nodal structures of the states are examined.
The calculated critical interaction strength $k_{F}a$ for ferromagnetic transition is shown to be
underestimated by a factor of two under the scattering length approximation.
Although we solved the problem
only for 4 particles, this minimal model suffices to show that ignoring the molecular states
with the scattering length approximation leads to
inaccurate results in the strongly interacting regime. Hence it leads to severe errors in many-body calculations.
That we get very good agreement with experimental estimates is encouraging but could be a result of cancellation of errors between the 4 particle system and the thermodynamic limit.
We investigated the dynamics of pair formation.
Non-monotic behavior of the pairing parameter $\sum_{i<j}\delta_{\mathbf{r}_{i},\mathbf{r}_{j}}$
is observed as a function of the final interaction strength $k_{F}a$ after a quench.
\textit{Acknowledgments}---This work has been supported
with funds from the DARPA OLE Program and ARO (Grant
no.~56693PH), computer time at NCSA. This work was initiated at
the Kavli Institute of Theoretical Physics in Beijing and Santa
Barbara.
\newpage
|
\section{Introduction}
Inflation \cite{Guth:1980zm,Linde:1981mu,Albrecht:1982wi} provides a compelling explanation for the large-scale homogeneity of the universe and for the observed spectrum of cosmic microwave background (CMB) anisotropies. However, in a large fraction of the multitude of inflationary models --- e.g., in most small-field models --- the success and the predictions of inflation are sensitive to small changes in the inflaton Lagrangian and initial conditions.
Without a priori measures on the space of scalar field Lagrangians and on the corresponding phase space, it is difficult to test a given model of inflation
In this paper we find robust predictions in a surprising place: warped D-brane inflation \cite{Kachru:2003sx},
a well-studied scenario for inflation in string theory in which six (or more) dynamical fields are governed by a scalar potential with hundreds of terms. We show that the collective effect of many terms in the potential is accurately described by a simple and predictive phenomenological model. The essential idea behind our approach is that in an inflationary model whose potential involves the sum of many terms depending on multiple fields, one can expect a degree of emergent simplicity, which may be thought of as central limit behavior.
Our primary method is a comprehensive Monte Carlo analysis. Recent results \cite{Baumann:2010sx} provide the structure of the scalar potential in warped D-brane inflation, i.e.\ a list of all possible terms in the potential, with undetermined, model-dependent coefficients. A realization of warped D-brane inflation then consists of a choice of coefficients together with a choice of initial conditions. We construct an ensemble of
realizations, drawing the coefficients from a range of statistical distributions and truncating the potential to contain $27$, $237$, and $334$ independent terms, corresponding to contributions from Planck-suppressed operators with maximum dimensions of $6$, $7$, and $\sqrt{28}-3/2 \approx 7.79$, respectively.
We then numerically evolve the equations of motion for the homogeneous background and identify robust observables that have demonstrably weak dependence on the statistical distribution, on the degree of truncation, and on the initial data. In particular, we find that the probability of $N_{e}$ e-folds of inflation is a power law, $P(N_{e})\propto N_{e}^{-3}$, and we present a very simple analytical model of inflection point inflation that reproduces this exponent.
To study the primordial perturbations, we focus on
the subset of realizations in which the dynamics during the final 60 e-folds is that of single-field slow roll inflation. (In the remaining realizations, multifield effects can be significant, and a dedicated analysis is required.) For these cases, we find that primordial perturbations consistent with WMAP7 \cite{Larson:2010gs} constraints on the scalar spectral index, $n_{s}$, are possible only in realizations yielding $N_{e} \gtrsim 120$ e-folds. In favorable regions of the parameter space, a universe consistent with observations arises approximately once in $10^5$ trials.
The plan of this paper is as follows. In \S\ref{review} we recall the setup of warped D-brane inflation, and in \S\ref{method} we explain how we construct and study an ensemble of realizations. Results for the homogeneous background evolution appear in \S\ref{background}, while the perturbations are studied in \S\ref{experiment}. We conclude in \S\ref{conclusions}. Appendix \ref{scalar potential} summarizes the structure of the inflaton potential, following \cite{Baumann:2010sx}.
\section{Review of Warped D-brane Inflation} \label{review}
In the simplest models of warped D-brane inflation,\footnote{See \cite{Dvali:1998pa}, \cite{Dvali:2001fw}, \cite{Burgess:2001fx} for foundational work on brane inflation.} the inflaton field $\phi$ is identified as the separation between a D3-brane and an anti-D3-brane along the radial direction of a warped throat region of a flux compactification \cite{Kachru:2003sx}. (We will have much to say about more complicated inflationary trajectories that involve angular motion.)
The D3-brane potential receives a rich array of contributions, from the Coulomb interaction of the brane-antibrane pair, from the coupling to four-dimensional scalar curvature, and from nonperturbative effects that stabilize the K\"ahler moduli of the compactification. The curvature coupling yields a significant inflaton mass,
and in the absence of any comparable contributions, the slow roll parameter $\eta$ obeys $\eta \approx 2/3$ \cite{Kachru:2003sx}, which is inconsistent with prolonged inflation.
The moduli-stabilizing potential does generically make significant contributions to the inflaton potential, and many authors have taken the attitude that within the vast space of string vacua, in some fraction the moduli potential will by chance provide an approximate cancellation of the inflaton mass, so that $\eta \ll 1$. To do better, one needs to know the form of the moduli potential. The nonperturbative superpotential was computed in \cite{Baumann:2006th} for a special class of configurations in which a stack of D7-branes falls inside the throat region. For this case, Refs.~\cite{Burgess:2006cb,Baumann:2007np,Krause:2007jk,Baumann:2007ah} studied the possibility of inflation, and found that fine-tuned inflation, at an approximate inflection point, is indeed possible \cite{Baumann:2007np,Krause:2007jk,Baumann:2007ah}.
This situation is unsatisfactory in several ways. First, the restriction to compactifications in which D7-branes enter the throat is artificial, and serves to enhance the role of known terms in the inflaton potential (those arising from interactions with the nearby D7-branes) over more general contributions from the bulk of the compactification. Second, the analyses of \cite{Burgess:2006cb,Baumann:2007np,Krause:2007jk,Baumann:2007ah,Panda:2007ie,Ali:2008ij,Chen:2008ai,Ali:2010jx} treated special cases in which the D3-brane tracked a minimum along some or all of the angular directions of the conifold, sharply reducing the dimensionality of the system, but there is no reason to believe that this situation is generic. Therefore, although these works do provide consistent treatments of inflation in special configurations, an analysis that studies the full six-dimensional dynamics in a general potential is strongly motivated.\footnote{See \cite{Chen:2008ada,Chen:2010qz} for detailed studies of multifield effects at the end of D-brane inflation in the framework of \cite{Baumann:2007ah}, and \cite{Hoi:2008gc} for a systematic exploration of the likelihood of inflation in this context.}
The results of \cite{Baumann:2008kq, Baumann:2010sx} provide the necessary information about the D3-brane potential. As explained in detail in \cite{Baumann:2010sx}, the most general potential for a D3-brane on the conifold corresponds to a general supergravity solution in a particular perturbation expansion around the Klebanov-Strassler solution. The most significant terms in this potential arise from supergravity modes corresponding to the most relevant operators in the dual CFT, and by consulting the known spectrum of Kaluza-Klein modes, one can write down the leading terms in the inflaton potential, up to undetermined Wilson coefficients. The physical picture is that effects in the bulk of the compactification, e.g. gaugino condensation on D7-branes, or distant supersymmetry breaking, distort the upper reaches of the throat, leading to perturbations of the solution near the location of the D3-brane.
To describe the D3-brane action, we begin with the background geometry, which is a finite region of the warped deformed conifold. Working far above the tip and ignoring logarithmic corrections to the warp factor, the line element is
\begin{equation}
ds^2 = \left(\frac{R}{r}\right)^2 g_{ij}dy^{i}dy^{j} = \left(\frac{R}{r}\right)^2 \Bigl(dr^2 +r^2 ds^2_{T^{1,1}}\Bigr)\,,
\end{equation} where $r$ is the radial direction of the cone, and the base space $T^{1,1}$ is parameterized by five angles, $0\le \theta_1,\theta_2 \le \pi$, $0 \le \phi_1,\phi_2 < 2\pi$, $0 \le \psi < 4\pi$, which we shall collectively denote by $\Psi$. The radius $R$ is given by $R^{4} =\frac{27}{4}\pi g_{s} N {\alpha^{\prime}}^2$, with $N \gg 1$ the D3-brane charge of the throat.
At radial coordinate $r_{\rm UV} \approx R$, the throat smoothly attaches to the remainder of the compact space, which we refer to as the bulk. On the other hand, the deformation is significant in the vicinity of the tip, at $r_0 \approx a_{0} R$, with $a_{0}$ denoting the warp factor at the tip. We will perform our analysis in the region $a_0 R \ll r < r_{\rm UV}$, where the singular conifold approximation is applicable, and will work with a rescaled radial coordinate $x\equiv \frac{r}{r_{\rm UV}} <1$.
Finally, because the D3-brane kinetic term is insensitive to warping at the two-derivative level (cf.\ \S\ref{dbi}), the metric on the inflaton field space is the {\it{unwarped}} metric $g_{ij}$.
The D3-brane potential is usefully divided into four parts,
\begin{equation} \label{thepotential}
V(x,\Psi)=V_0+V_{\cal{C}}(x) + V_{{\cal{R}}}(x) + V_{\rm bulk}(x,\Psi)\,,
\end{equation}
which we will discuss in turn.
First, the constant $V_0$ represents possible contributions from distant sources of supersymmetry breaking, e.g. in other throats.
Next, the Coulomb potential $V_{\cal{C}}$ between an anti-D3-brane at the bottom of the throat and the mobile D3-brane has the leading terms
\begin{equation}
V_{\cal{C}}=D_{0}\left(1-\frac{27D_{0}}{64\pi^2 T_{3}^2 r_{\rm UV}^4}\frac{1}{x^{4}}\right)\,,
\end{equation}
where $T_3$ is the D3-brane tension and $D_{0}=2a_0^4 T_3$. Higher-multipole terms in the Coulomb potential depend on the angles $\Psi$, but are suppressed by additional powers of $a_0$ and may be neglected in our analysis.
Upon defining the scale $\mu^4=(V_0+D_0)\left(\frac{T_{3}r_{\rm UV}^{2}}{M_{\rm pl}^{2}}\right)$, the leading contribution from curvature, corresponding to a conformal coupling, may be written
\begin{equation}
V_{{\cal{R}}}= \frac{1}{3}\mu^4 x^2\,.
\end{equation}
Finally, the structure of the remaining terms has been obtained in \cite{Baumann:2010sx}:
\begin{equation} \label{thepotential2}
V_{\rm bulk}(x,\Psi)= \mu^4 \sum_{LM} c_{LM} x^{\delta(L)} f_{LM}(\Psi)\,.
\end{equation}
Here $LM$ are multi-indices encoding the quantum numbers under the $SU(2) \times SU(2) \times U(1)$ isometries of $T^{1,1}$, the functions $f_{LM}(\Psi)$ are angular harmonics on $T^{1,1}$, and $c_{LM}$ are constant coefficients.
The exponents $\delta(L)$ have been computed in detail in \cite{Baumann:2010sx}, building on the computation of Kaluza-Klein masses in \cite{Ceresole}:
\begin{equation}
\delta=1,\, 3/2,\, 2,\,\sqrt{28}-3,\, 5/2,\,\sqrt{28}-5/2,\,3,\,\sqrt{28}-2,\,7/2,\,\sqrt{28}-3/2, \ldots
\end{equation}
From the viewpoint of the low-energy effective field theory, a term in the potential proportional to $x^{\delta(L)}$ arises from a Planck-suppressed operator with dimension $\Delta=\delta(L)+4$. In particular, the conformal coupling to curvature corresponds to an operator of dimension six, ${\cal O}_{6} =(V_0+D_0)\phi^2$.
Two technical remarks are in order. First, for simplicity of presentation we have included higher-order curvature contributions in the list of bulk terms, rather than in a separate category. Second, perturbations of the unwarped metric $g_{ij}$ lead to terms in the D3-brane potential that were not analyzed in \cite{Baumann:2010sx}, but can be important in some circumstances.\footnote{We thank Sohang Gandhi for very helpful discussions of this point.} We do not implement the angular structure of these terms in full detail, but we have verified that these contributions lead to negligible corrections to our results.
The coefficients $c_{LM}$ could be computed in principle in a specific realization in which all details of the compactification are available, but in practice must be treated as unknown parameters.
Our approach is to assume that all possible terms are present, with coefficients $c_{LM}$ of comparable magnitude. Specifically, we will draw the $c_{LM}$ from a range of statistical distributions and then verify that the
(unknown) detailed statistical properties of the $c_{LM}$ are not important for the inflationary phenomenology, while the overall scale of the $c_{LM}$ does matter significantly.
We set the overall scale of the bulk contributions by noting that the moduli potential and the remainder of the potential are tied by the requirement that the cosmological constant should be small after brane-antibrane annihilation. With our definition of $\mu$, the scaling arguments presented in Appendix A of \cite{Baumann:2008kq} suggest that in typical KKLT compactifications, $c_{LM} \sim {\cal{O}}(1)$.
Let us remark that the potential (\ref{thepotential}) is not the most general function on the conifold: it is the most general D3-brane scalar potential on the conifold (within the fairly broad assumptions of \cite{Baumann:2010sx}.) Many terms in (\ref{thepotential}) enjoy correlations that would be absent in a totally general function, and which arise here because certain physical sources, such as fluxes, contribute in correlated ways to different terms in the potential. We defer a full description of the construction of the potential to Appendix \ref{scalar potential}.
Finally, we note that in compactifications preserving discrete symmetries that act nontrivially on the throat region, the structure of the D3-brane potential is altered by the exclusion of terms that are odd under the discrete symmetries \cite{Baumann:2008kq}. Exploring the phenomenology of the corresponding models is an interesting question that is beyond the scope of this work.\footnote{We thank Daniel Baumann for helpful discussions of this point.}
\section{Methodology} \label{method}
To characterize the dynamics of D3-brane inflation in a general potential, we perform a Monte Carlo analysis, numerically evolving more than $7 \cdot 10^7$ distinct realizations of the model.
In this section we explain our recipe for constructing an ensemble of realizations. In \S\ref{evolution}, we obtain the
equations of motion and introduce the parameters required to specify the potential. In \S\ref{parameters} we describe how we draw the coefficients in the potential from statistical distributions, and in
in \S\ref{rotational} we indicate how we choose initial conditions.
\subsection{Setup}\label{evolution}
The inflaton field is characterized by one radial coordinate and five angular coordinates. At the two-derivative level (see \S\ref{dbi} for a discussion of DBI effects), the equations of motion for the homogeneous background are the Klein-Gordon equations obtained from the Lagrangian
\begin{equation}
\mathcal{L} = a^{3}\left(\frac{1}{2}T_3 g_{ij}\dot{y}^{i}\dot{y}^{j} - V(y) \right)\,,
\end{equation}
where $a$ is the scale factor, along with the Friedmann and acceleration equations,
\begin{eqnarray}
3H^{2} & = & \frac{1}{2}T_3 g_{ij}\dot{y}^{i}\dot{y}^{j} + V(y)\,, \\
\dot{H} & = & -\frac{1}{2}T_3 g_{ij}\dot{y}^{i}\dot{y}^{j}\,.
\end{eqnarray}
Here
$y^{i}$ denotes the six coordinates, dots indicate derivatives with respect to time, $H \equiv \dot{a}/a$, and $g_{ij}$ is the metric on the inflaton field space, which is the conifold.
Three important microphysical parameters are the D3-brane tension, $T_3$; the length of the throat, $\phi_{\rm UV} \equiv r_{\rm UV} \sqrt{T_{3}}$; and the warp factor at the tip, $a_{0}$. The combination $2 T_3 a_{0}^{4} \equiv D_{0}$ determines the overall scale of inflation, while $\phi_{\rm UV}$ dictates the size of the field space.
The result of \cite{Baumann:2006cd} gives an upper bound on the inflaton field range, $\phi_{\rm UV}<\frac{2 M_{\rm pl}}{\sqrt{N}}$, with $N \gg 1$ the D3-brane charge of the warped throat. Consistent with this, we take $\phi_{\rm UV}=0.1$. Working in units where $M_{\rm pl}^{-2}=8\pi G =1$ for the remainder, we set the D3-brane tension to be $T_3=10^{-2}$, and for our analysis of the background evolution, we take $a_0=10^{-3}$.
We have verified that changing these parameters does not substantially alter our results
for the homogeneous background. However, changing $T_3 a_{0}^{4}$ --- which we accomplish by changing $a_{0}$ --- does affect the scale of inflation, and hence the normalization of the scalar perturbations. Therefore, in our study of the perturbations in \S\ref{experiment}, we scan over a range of values for $a_0$, focusing on values most likely to lead to a WMAP-normalized spectrum \cite{Larson:2010gs}. This is a fine-tuning that we will not attempt to quantify, as there is no agreed-upon measure for $a_0$.
\subsection{Constructing an ensemble of potentials} \label{parameters}
In principle the D3-brane potential (\ref{thepotential}) has an infinite number of terms,
but for $x\equiv \frac{r}{r_{\rm UV}} < 1$ one can truncate (\ref{thepotential}) at some maximum exponent $\delta=\delta_{\rm max} \equiv \Delta_{\rm max} -4$.
Because of the critical role of the inflaton mass term, truncating to $\Delta_{\rm max}<6$ would fail to capture essential physical properties, so we must have $\Delta_{\rm max} \geq 6$. As $\Delta$ increases, the number of independent terms grows very rapidly, because there are
many angular harmonics $f_{LM}$ for each $\Delta$.
Limited by computational power, we truncate the potential at $\Delta_{\rm max}= \sqrt{28}+5/2 \approx 7.8$. We perform identical analyses for $\Delta_{\rm max}=6$, $7$, and $7.8$, corresponding respectively to 27, 237 and 334 independent terms in the potential, in order to assess whether our results are sensitive to the cutoff.
Many studies of D-brane inflation treat the evolution of the radial position, the volume of a particular four-cycle, and sometimes one angular coordinate, cf.\ e.g.\ \cite{Burgess:2006cb,Baumann:2007np,Krause:2007jk,Baumann:2007ah,Panda:2007ie,Ali:2008ij,Chen:2008ai,Ali:2010jx}, rather than the full multifield dynamics. Although angular evolution in the framework of \cite{Baumann:2007ah} has been studied in detail in \cite{Chen:2008ada,Chen:2010qz}, the focus of these works was the onset of angular instabilities at the end of inflation. We will find that angular evolution before the onset of inflation also plays a critical role.
To understand how our results differ from treatments with fewer dynamical fields, we study the impact of stepwise increases in the number $N_f$ of evolving fields. We artificially, but self-consistently, freeze $6-N_f$ of the angular fields by not imposing the corresponding equations of motion, creating realizations that depend on $N_f$ variables. While these realizations have less physical meaning than the full potential, they provide some insight into the role of the angular fields. For simplicity we study $N_f=1$, $N_f=2$, and the full case $N_{f}=6$.
As we are not assuming that D7-branes wrapping a four-cycle descend into the throat region, we will not model the evolution of the K\"ahler moduli. Although it would be very interesting (and challenging)
to study the cosmological dynamics of K\"ahler moduli in the bulk, the universality found in the present analysis makes it plausible that additional fields would have little effect on the inflationary phenomenology.
Turning now to the Wilson coefficients $c_{LM}$, we do not assume a specific compactification, but instead draw the $c_{LM}$ from a range of statistical distributions.
We define the root mean square (rms) size, $\langle c_{LM}^2 \rangle^{1/2} \equiv Q$, where the brackets denote the ensemble average, and by assumption\footnote{A strong trend in $Q$ as a function of $\Delta$ could change the relative importance of terms with large $\Delta$, and hence affect our conclusions about the robustness of the truncation to $\Delta \le \Delta_{\rm max}$. We are not aware of a well-motivated proposal for such a trend, but it could be worthwhile to investigate this further.} the rms size is independent of $L$ and $M$.
It is then convenient to write
\begin{equation}\label{qhat}
c_{LM} = Q \,\hat{c}_{LM}\,,
\end{equation} and draw the $\hat{c}_{LM}$ from some distribution ${\cal M}$ that has unit variance but is otherwise arbitrary.
The physical picture is that $Q$ depends on the distance to the nearest stack of D7-branes effecting K\"ahler moduli stabilization. The estimates performed in \cite{Baumann:2008kq} indicate that $Q \sim {\cal{O}}(1)$ for
D7-branes in the upper region of the throat. We anticipate that as the nearest D7-branes are moved farther into the bulk, $Q$ will diminish to some extent, though we are not aware of a regime in which the bulk contributions are strictly negligible.
If the inflationary phenomenology depended in detail on the nature of ${\cal M}$, e.g. if the success of inflation depended sensitively on the higher moments of ${\cal M}$, then no general predictions would be possible. Let us clarify that dependence on the rms size $Q$ of the $c_{LM}$, corresponding to the typical size of the bulk contribution to the inflaton potential, is to be expected and is not problematic. Difficulty would arise if, for example, two distributions with unit variance but with distinct skewness or kurtosis led to disparate predictions.
There are strong motivations for expecting that some statistical properties of the potential will be independent of ${\cal M}$. For example, if a symmetric $N \times N$ matrix has its
entries drawn from some distribution with appropriately bounded moments, then in the large $N$ limit the statistical properties of the eigenvalues are indistinguishable from those obtained from entries drawn from a Gaussian distribution with mean zero \cite{Bai}.
By experimenting with different distributions, we will identify observables which, like the eigenvalue distribution in random matrix theory, are robust against changes in the statistics of the inputs.
In practice, for much of our analysis we choose ${\cal M}$ to be a Gaussian distribution with mean zero,
and then carefully verify for a range of other distributions that our results receive negligible corrections.
\subsection{Initial conditions}\label{rotational}
The phase space of initial conditions for a D3-brane in the conifold is 12-dimensional: six dimensions for the initial positions $x_{0},\Psi_{0}$ and another six dimensions for the initial velocities $\dot{x}_{0}, \dot{\Psi}_{0}$. A grid-based scan across the full 12-dimensional space would be very computationally intensive even with only a few points along each dimension. Fortunately, five of the six dimensions are angular coordinates on the coset space $T^{1,1}$, which has a large isometry group, $SU(2) \times SU(2) \times U(1)$. These isometries can be used to reduce the dimensionality of the initial phase space, in the following way.
A generic configuration of sources in the compact space will break the isometry group completely, but in a large ensemble of realizations, we expect that there are no preferred regions on $T^{1,1}$: the ensemble averages should respect the isometries even though any individual realization breaks the isometries.
Thus, without loss of generality we may pick a fixed point $\Psi_{0}$ on $T^{1,1}$ for the initial position. For numerical purposes it is convenient to begin away from the coordinate singularities, so we choose $\Psi_{0}$ to be $\theta_1=\theta_2=\phi_1=\phi_2=\psi=1.0$.
The initial angular velocities $\dot{\Psi}_{0}$ are slightly more complicated.\footnote{We are grateful to Raphael Flauger for helpful discussions of this point.} To describe a general angular velocity, it suffices to specify the magnitude of the velocity in each $S^2$ and in the fiber $S^{1}$. For simplicity we focus on velocity in the fiber, $\dot{\psi}$, and take the remaining components of the initial velocity to vanish.
We expect, and find, similar results for initial velocities in either $S^2$, but we postpone a complete scan of the phase space to future work.
We are left with a three-dimensional space of initial configurations spanned by the radial position $x_{0}$, the radial velocity $\dot{x}_{0}$, and the angular velocity $\dot{\Psi}_{0}=\dot{\psi}_{0}$. Of course, our evolution occurs in the full 12-dimensional phase space: the simplification applies only to the initial conditions. For a portion of our Monte Carlo analysis, we set $\dot{x}_{0}=\dot{\psi}_{0}=0$, so that the D3-brane begins at rest. In \S\ref{iniv} we describe the effect of nonvanishing initial velocities.
\subsection{Parameters summarized}
To summarize, we fix $T_3=10^{-2}$, $a_0=10^{-3}$, and $\phi_{\rm UV}=0.1$ for our analysis of the background evolution.
We truncate the D3-brane potential to include contributions from operators with maximum dimension $\Delta_{\rm{max}}=6$, $7$, and $7.8$, and we take $N_f=1,2,6$ of the D3-brane coordinates to be dynamical fields. The coefficients $c_{LM}$ have rms size $Q$, and the rescaled quantities $\hat{c}_{LM} = c_{LM}/Q$ are drawn from a distribution ${\cal M}$ that has unit variance.
We begin at $x=x_{0}\equiv 0.9$, $\Psi=\Psi_{0} \equiv \{\theta_1=\theta_2=\phi_1=\phi_2=\psi=1.0\}$, with arbitrary radial velocity $\dot{x}_{0}$, arbitrary angular velocity $\dot{\psi}_{0}$ in the $\psi$ direction, and all other angular velocities vanishing.
We would now like to understand how the observables depend on the input parameters $Q$, $\Delta_{\rm{max}}$, $N_f$, and ${\cal M}$, and on the initial data $x_{0},\dot{x}_{0},\dot{\psi}_{0}$.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.0in,angle=0]{Bouquet1.png}
\caption{Examples of downward-spiraling trajectories for a particular realization of the potential. The black dots mark 60 and 120 e-folds before the end of inflation (7 of the 8 curves shown achieve $N_e>120$); inflation occurs along an inflection point that is not necessarily parallel to the radial direction. Red curves have nonvanishing initial angular velocities $\dot{\Psi}_{0}$, while blue curves have $\dot{\Psi}_{0}=0$.}
\label{bouquet}
\end{center}
\end{figure}
\section{Results for the Homogeneous Background}
\label{background}
As a first step, we study the evolution of the homogeneous background. In \S\ref{probability}, we show that for fixed initial conditions, the probability of $N_e$ e-folds of inflation is a power law, and we show that the exponent is robust against changes in the input parameters $\Delta_{\rm max}$, $N_f$, and ${\cal{M}}$.
In \S\ref{analyticsection} we present a simple analytic model that reproduces this power law. We study the effect of varying the initial conditions in \S\ref{initial}, and we discuss DBI inflation in \S\ref{dbi}.
\subsection{The probability of inflation}
\label{probability}
We find it useful to divide possible trajectories into three classes. The D3-brane can be ejected from the throat, reaching $x>1$ and leaving the domain of validity of our analysis; it can become trapped in a local or global minimum of the potential; and it can reach the bottom\footnote{In practice, we define the bottom of the throat to be at $x = 20 a_0$ in order to remain well above the region where the throat rounds off and the singular conifold approximation fails.}
of the throat, triggering the hybrid exit and reheating, after a certain number of e-folds of inflation.
A central question is what fraction of realizations solve the horizon problem by producing $N_{e} \ge 60$ e-folds of inflation, and then plausibly transition to the hot Big Bang. We will not model reheating in detail, but we will insist that only e-folds of inflation {\it{that precede a hybrid exit}} are counted towards $N_{e}$. That is, false vacuum inflation in a metastable minimum, or slow roll inflation preceded by ejection and unknown dynamics in the bulk, do not contribute to $N_{e}$.
Specifically, for $k$ distinct trials we define
\begin{equation}
P(N_e>60) = \frac{\#(N_e>60)}{\#(N_e>60)+\#(N_e\le60)+\#(\rm{ejected})+\#(\rm{trapped})} \,,
\end{equation}
i.e.\ trials leading to ejection or trapping are included in the denominator, so that $P(N_e>60)$ reflects the probability of $N_{e} > 60$ e-folds of inflation preceding a hybrid exit
in a general realization.
We now examine how the `success probability' $P(N_e>60)$ depends on the input parameters $Q$, $\Delta_{\rm max}$, $N_f$, and ${\cal{M}}$. First, Figure \ref{fig:Q_histogram} shows that $P(N_e>60)$ depends strongly on
$Q$, and
the optimal value of $Q$ depends on $\Delta_{\rm max}$ and on $N_f$.
When all six fields are dynamical ($N_f=6$), the probability of inflation is optimized for $Q\sim 0.04$, while for $N_f=1$, $Q\sim1$ can yield sufficient inflation.
To understand this result, we recall that in the presence of a single harmonic contribution to the inflaton potential, after minimization of the angular potential, the radial potential is expulsive \cite{Baumann:2008kq}. More generally, the bulk contributions to the potential provide the only possibility of counterbalancing the Coulomb and curvature contributions, which both draw the D3-brane towards the tip. For $Q=0$, the Coulomb and curvature contributions are not counterbalanced, and the D3-brane falls quickly towards the tip without driving inflation. For $Q \sim 1$, a single harmonic contribution term could marginally balance the inward force; the net effect of 334 such terms then plausibly leads to rapid expulsion from the throat. This result is consistent with our finding that the optimal value of $Q$ diminishes as the number of terms in the potential increases, as shown in Figure \ref{fig:Q_histogram}.
\begin{figure*}[!h]
\begin{center}
\mbox{
{\includegraphics[width=3.in,angle=0]{Q_histograms_1.pdf}}\quad
{ \includegraphics[width=3.in,angle=0]{Q_histograms_2.pdf}}
}
\caption{The rms value, $Q$, of the coefficients $c_{LM}$ has a significant role in determining whether inflation can occur.
[Left panel] The success probability $P(N_e>60)$ for two different numbers of fields, $N_f=1$ and $N_f=6$, with $\Delta_{\rm max}=7.8$. [Right panel] The success probability for two different degrees of truncation, $\Delta_{\rm max}=6$ and $\Delta_{\rm max}=7.8$, with $N_f=6$.}
\label{fig:Q_histogram}
\end{center}
\end{figure*}
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.in,angle=0]{gamma_Q_1.pdf}
\includegraphics[width=3.in,angle=0]{gamma_Q_2.pdf}
\caption{The likelihood of $N_{e}$ e-folds of inflation as a function of $N_{e}$, for $\Delta_{\rm max} = 7.8$ and $N_{f} = 6$.
We find $P(N_e)\propto N_e^{-\alpha}$, with $\alpha=3.22\pm 0.07$ at the 68\% confidence level. The left panel shows the power law fit to the histogram and the right panel shows the same fit on a log-log plot.}
\label{fig:gamma_histogram}
\end{center}
\end{figure}
In Figure \ref{fig:gamma_histogram} we display a histogram of Monte Carlo trials that give more than 40 e-folds of inflation for $Q \in [0,2.0]$ with $\Delta_{\rm max}=7.8$ and all six fields evolving ($N_{f} = 6$). We find that we can characterize the probability of inflation, for scenarios yielding $N_{e} \gg 10$ e-folds, by a function $P(N_{e}) = {\cal{A}}\, (N_{e}/60)^{-\alpha}$. On fitting the data in Figure \ref{fig:gamma_histogram}, we find that $\alpha = 3.22 \pm 0.07$ and ${\cal{A}} = 1.7 \cdot 10^{-6}$.
\begin{table}[t!]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
$N_f$ & $\Delta_{\rm max}$ & ${\mathcal A}$ & $\alpha$ & $P(N_{e}>60)$ & $P(N_{e}>120)$ & $P({\rm ej})$ & $P({\rm min})$
\\
\hline
6 & 7.8 & $1.7 \cdot 10^{-6}$ & $3.22 \pm 0.07$ & $4.6 \cdot 10^{-5}$ & $1.1 \cdot 10^{-5}$ & $0.97$ & $3.9 \cdot 10^{-3}$
\\
6 & 7 & $3.2 \cdot 10^{-6}$ & $3.19 \pm 0.10$ & $9.0 \cdot 10^{-5}$ & $2.0 \cdot 10^{-5}$ & $0.96$ & $8.8 \cdot 10^{-3}$
\\
6 & 6 & $4.3 \cdot 10^{-6}$ & $2.85 \pm 0.14$ & $1.3 \cdot 10^{-4}$ & $3.3 \cdot 10^{-5}$ & $0.92$ & $5.4 \cdot 10^{-2}$
\\
\hline
2 & 7.8 & $3.0 \cdot 10^{-5}$ & $3.30 \pm 0.25$ & $9.6 \cdot 10^{-5}$ & $2.4 \cdot 10^{-5}$ & $0.77$ & $6.7 \cdot 10^{-2}$
\\
2 & 7 & $5.6 \cdot 10^{-5}$ & $2.97 \pm 0.17$ & $1.8 \cdot 10^{-4}$ & $4.9 \cdot 10^{-5}$ & $0.66$ & $1.2 \cdot 10^{-1}$
\\
2 & 6 & $1.5 \cdot 10^{-5}$ & $2.99 \pm 0.12$ & $4.7 \cdot 10^{-4}$ & $1.3 \cdot 10^{-4}$ & $0.51$& $1.9 \cdot 10^{-1}$
\\
\hline
1 & 7.8 & $2.7 \cdot 10^{-6}$ & $2.83 \pm 0.22$ & $8.8 \cdot 10^{-5}$ & $2.6 \cdot 10^{-5}$ & $0.36$ & $5.1 \cdot 10^{-2}$
\\
1 & 7 & $4.3 \cdot 10^{-6}$ & $3.03 \pm 0.21$ & $1.2 \cdot 10^{-4}$ & $2.9 \cdot 10^{-5}$ & $0.29$ & $6.1 \cdot 10^{-2}$
\\
1 & 6 & $1.3 \cdot 10^{-5}$ & $2.86 \pm 0.15$ & $4.2 \cdot 10^{-4}$ & $1.1 \cdot 10^{-4}$ &$0.17$ &$7.4 \cdot 10^{-2}$
\\
\hline
\end{tabular}
\caption{Summary of the power law fits to the success probabilities for various scenarios, with $Q$ varied over the range $[0.0,2.0]$. We fit the Monte Carlo data for $N_e>40$ to a power law of the form $P(N_e) = {\cal{A}}\,(N_e/60)^{-\alpha}$. The two final columns indicate the probabilities that the D3-brane will be ejected from the throat or trapped in a minimum, respectively.
}
\label{table:gamma}
\end{center}
\end{table}
In Table \ref{table:gamma} we summarize the power law fits to the probability of inflation as one considers different numbers of fields, $N_f$, and different truncations of the potential, $\Delta_{\rm max}$, assuming that ${\cal M}$ is Gaussian, and taking zero initial angular and radial velocities at $x=0.9$ and fixed angular position $\Psi_{0}$.
The probability of obtaining 60 e-folds of inflation does not change dramatically if one truncates the potential at $\Delta_{\rm max} = 6.0$ or $\Delta_{\rm max} = 7.8$, so our results appear insensitive to the precise placement of the truncation. As the number of fields, $N_f$, increases, the range of $Q$ yielding inflation becomes restricted, but we also find that the probability of achieving inflation within this $Q$ range increases, cf.\ Figure \ref{fig:Q_histogram}. In fact, the power law fit of the success probability remains fairly consistent as $N_f$ is varied, provided that one marginalizes over $Q$.
Although we have seen that $P(N_e>60)$ is highly sensitive to the value of $Q$, we find that $P(N_e>60)$ has negligible dependence on the shape of the distribution ${\cal M}$ from which the $\hat{c}_{LM}$ are drawn.
Specifically, we have obtained power law fits of the success probability for ensembles in which ${\cal M}$ is a Gaussian, shifted Gaussian, triangular, or uniform distribution.
As shown in Table \ref{table:measure}, we find negligible changes in ${\cal A}$ and $\alpha$.
We expect that there exist pathological distributions, e.g.\ with rapidly growing higher moments, that could change our findings, but we are not aware of a microphysical argument for such a distribution.
\begin{table}[t!]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
${\cal M}$ & $\Delta_{\rm max}$ & ${\mathcal A}$ & $\alpha$ & $P(N_{e}>60)$ & $P(N_{e}>120)$
\\
\hline
Gaussian& 7.8 & $1.6 \cdot 10^{-6}$ & $3.21 \pm 0.14$ & $4.6 \cdot 10^{-5}$ & $1.1 \cdot 10^{-5}$
\\
shifted & 7.8 & $1.7 \cdot 10^{-6}$ & $3.42 \pm 0.10$ & $4.7 \cdot 10^{-5}$ & $1.1 \cdot 10^{-5}$
\\
triangular & 7.8 & $1.8 \cdot 10^{-6}$ & $3.21 \pm 0.09$ & $4.9 \cdot 10^{-5}$ & $1.1 \cdot 10^{-5}$
\\
uniform & 7.8 & $1.7 \cdot 10^{-6}$ & $3.18 \pm 0.08$ & $4.8 \cdot 10^{-5}$ & $1.1 \cdot 10^{-5}$
\\
\hline
Gaussian & 6 & $4.3 \cdot 10^{-6}$ & $2.96 \pm 0.09 $ & $1.3 \cdot 10^{-4}$ & $3.4 \cdot 10^{-5}$
\\
shifted& 6 & $3.7 \cdot 10^{-6}$ & $3.03 \pm 0.08 $ & $1.1 \cdot 10^{-4}$ & $2.9\cdot 10^{-5}$
\\
triangular & 6 & $4.3 \cdot 10^{-6}$ & $2.94 \pm 0.08$ & $1.3 \cdot 10^{-4}$ & $3.5 \cdot 10^{-5}$
\\
uniform & 6 & $4.3 \cdot 10^{-6}$ & $3.06 \pm 0.07$ & $1.3 \cdot 10^{-4}$ & $3.6 \cdot 10^{-5}$
\\
\hline
\end{tabular}
\caption{Summary of the power law fits to the success probabilities for various distributions ${\cal M}$,
with $Q$ varied over the range $[0.0,2.0]$, and for initial positions chosen randomly on $T^{1,1}$.
We fit the Monte Carlo data for $N_e>40$ to a power law of the form $P(N_e) = {\cal{A}}\,(N_e/60)^{-\alpha}$.}
\label{table:measure}
\end{center}
\end{table}
\subsection{An analytic explanation of the exponent $\alpha=3$}\label{analyticsection}
In our ensemble of potentials, inflation typically occurs near an approximate inflection point of the potential. We now show that a very simple model of single-field inflection point inflation, along the lines proposed in \cite{Freivogel:2005vv}, predicts $\alpha=3$, in excellent agreement with our numerical results.\footnote{We thank G. Shiu and H. Tye for very helpful discussions of this point.}
An approximate inflection point of a function $V(\phi)$ of a single field $\phi$ is a location where $V^{\prime\prime}=0$ and $V^{\prime}$ is small in appropriate units. We choose the origin of $\phi$ to correspond to the zero of $V^{\prime\prime}$, so that
\begin{equation} \label{inflectionmodel}
V(\phi) = c_{0} + c_{1} \phi + c_{3} \phi^3 +
\ldots\,,
\end{equation}
with the $c_{i}$ being constants. Assuming that the constant term dominates, the number of e-folds of inflation is
\begin{equation}
N_{e} \approx \frac{c_{0}}{\sqrt{c_{1}c_{3}}}
\end{equation}
in the regime of interest where the $c_{i}$ are small.
The approach suggested in \cite{Freivogel:2005vv} is to obtain the probability of $N_{e}$ e-folds of inflation by computing
\begin{equation} \label{unfixedpower}
P(N_{e}) = \int \prod_{i=1}^{k} d\xi_{i} F(\xi_{1}, \ldots \xi_{k})
\delta\Bigl(N_{e} - f(\xi_{1}, \ldots \xi_{k})\Bigr)\,,
\end{equation}
where the $\xi_{i}$ are the parameters of the model, $f$ is the number of e-folds as a function of these parameters, and $F$ is a measure on the parameter space. Determining $F$ from first principles is very subtle, and is beyond the scope of this work. However, to compare to our numerical results involving {\it{relative}} probabilities of different numbers of e-folds, we need only use a measure $F$ that properly represents the measure ${\cal{M}}$ that we have imposed on the coefficients in our ensemble. At very small values of the $c_{i}$, we can approximate ${\cal{M}}$ as a constant, and so we take $F(c_{0},c_{1},c_{3}) =1$. Thus, we need to evaluate
\begin{equation}
P(N_{e}) = \int d c_{1}\, d c_{3}\, \delta\left(N_{e}
- \frac{c_{0}}{\sqrt{c_{1}c_{3}}}\right)\,,
\end{equation}
Performing the integral and again using the smallness of the $c_{i}$, we find
\begin{equation}
P(N_{e}) \approx -\frac{4c_{0}^{2}}{N_{e}^{3}}{\rm{log}}(c_0)
\end{equation}
so that $\alpha=3$,
which compares very well to our numerical results displayed in Table \ref{table:gamma}.
In the homogeneous background analysis described in \S\ref{probability}, the power in scalar perturbations is unconstrained. However, in \S\ref{experiment} we will assemble realizations whose scalar perturbations are consistent with the WMAP7 \cite{Larson:2010gs} normalization. To compare to the ensemble of \S\ref{experiment} with fixed scalar power, we must compute
\begin{equation} \label{fixedpower}
P(N_{e}) = \int \prod_{i=1}^{k} d\xi_{i} F(\xi_{1}, \ldots \xi_{k})
\delta\Bigl(N_{e} - f(\xi_{1}, \ldots \xi_{k})\Bigr)\delta\Bigl(A_{s}^{\star} - A_{s}(\xi_{1}, \ldots \xi_{k})\Bigr)\,,
\end{equation} where $A_{s}(\xi_{1}, \ldots \xi_{k})$ is the amplitude of the scalar perturbations as a function of the parameters $\xi_{i}$, and $A_{s}^{\star}$ is the central value measured by WMAP7.
For the inflection point model (\ref{inflectionmodel}), when the scalar power is fixed as in (\ref{fixedpower}), one again finds $\alpha = 3$, just as in the case (\ref{unfixedpower}) with unconstrained scalar power.
Moreover, the ensemble of \S\ref{experiment} with fixed scalar power is consistent with $\alpha=3$, providing a second check of our analytical model.
\subsection{Dependence on initial conditions}
\label{initial}
By construction, there is no preferred angular position selected by the ensemble of potentials:
upon averaging over all possible source locations in the bulk,
we recover ensemble average rotational invariance. However, in any particular realization, the potential will be quite different at different angular locations, so it is meaningful to ask about the effect of varying the initial angular position in a given realization. Moreover, changes in the initial radial position can significantly alter the dynamics. In \S\ref{inip} we determine the effects of altering the initial radial and angular positions, while the effects of varying the initial velocities are presented in \S\ref{iniv}.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=4.0in,angle=0]{attractorplotJan303D_Liam.png}
\caption{Trajectories in the $\theta_{1}-\theta_{2}$ plane, with $\rm{log}(x)$ vertical, for a fixed potential. Notice the attractor behavior in the angular directions. Green trajectories correspond to $\approx 5$ e-folds of expansion, while the remaining colors correspond to trajectories with $\approx 150$ e-folds.}
\label{fig:attractor}
\end{center}
\end{figure}
\subsubsection{Dependence on the initial position} \label{inip}
Prior works on initial conditions for D-brane inflation have found that in many examples, the inflaton needs to begin with small velocity just above the inflection point in order to yield substantial inflation.
In our ensemble, overshooting is not a problem: in most realizations yielding at least 60 e-folds, it suffices to begin the evolution with small velocity high up in the throat, e.g. at $x = 0.9$, while the inflection point is generally in the vicinity of $x=0.1$ or even smaller. In fact, increasing the initial radial position typically {\it{increases}} the amount of inflation. We suggest that this increase could be due to an increased opportunity to find the inflection point during a prolonged period of radial infall.
The amelioration of the overshoot problem in our ensemble is a reflection of the difference between potentials that are fine-tuned by hand and potentials that are chosen randomly. In the former case, there is a natural tendency to fine-tune the potential to be just flat enough for 60 e-folds of inflation given perfect initial conditions, but no flatter. In contrast, when scanning through the space of possible potentials, one can actually find more robust examples. As we have seen, successful realizations are reasonably common.
The success of inflation has very mild dependence on the initial angular positions: we find that in realizations of the potential that yield more than 60 e-folds of inflation for one set of initial positions,
an order-unity fraction of the space of initial angular positions leads to the same outcome. Indeed, we find attractor behavior in the space of initial angles, as illustrated in Figure \ref{fig:attractor}.
\subsubsection{Dependence on the initial velocity}\label{iniv}
When the D3-brane begins with a radial velocity of order $10^{-6}$ of the local limiting speed (cf. \S\ref{dbi}), corresponding to an initial kinetic energy that is $\approx 10\%$ of the potential energy, it strikes the bottom of the throat within a fraction of an e-fold. However, initial {\it{angular}} velocity of the same magnitude has a different effect: the D3-brane is quickly ejected from the throat.
We have found that two distinct causes contribute
to this ejection effect: first, D3-branes with large angular velocities can overcome potential barriers in the angular directions and thereby explore a larger fraction of $T^{1,1}$, including regions where the potential is strongly expulsive. Second, angular momentum produces a barrier to radial infall, as in standard central force problems. Inward-directed radial velocity and comparably large angular velocity have counterbalancing effects
in many cases, suggesting that slow roll inflation could arise in special regions of phase space where the initial velocities are not small, but have compensating effects. We leave this as an interesting question for future work.
\subsection{The DBI effect}
\label{dbi}
In certain parameter regimes, higher-derivative contributions to the D3-brane kinetic energy can support a phase of DBI inflation \cite{Silverstein:2003hf,Alishahiha:2004eh}. The
DBI Lagrangian is
\begin{equation}
{\mathcal L} = a^{3}\left( -T(y) \sqrt{1 - \frac{T_{3}g_{ij}\dot{y}^{i}\dot{y}^{j}}{T(y)}} - V(y) + T(y) \right),
\end{equation}
where in the $AdS_{5}\times T^{1,1}$ approximation with warp factor $e^{A}=x$, $T(y) = T_{3}x^{4}$. DBI inflation can occur if the D3-brane velocity approaches the local limiting speed, i.e. if
\begin{equation}
1 - \frac{T_{3}g_{ij}\dot{y}^{i}\dot{y}^{j}}{T(y)} \equiv \frac{1}{\gamma^2} \rightarrow 0 \,.
\end{equation}
In our Monte Carlo trials,
we did not
observe a single example with $\gamma-1 > 10^{-8}$, so the DBI effect was never relevant in our system.
To understand this result, we recall that steep potentials are generically required to accelerate D3-branes to approach the local speed of light, and not every potential that is too steep to support slow roll inflation is actually steep enough to drive DBI inflation in a given warped background.\footnote{We thank Enrico Pajer for instructive discussions of this issue.} Specifically, DBI inflation requires \cite{Alishahiha:2004eh}
\begin{eqnarray} \label{steep}
\left(\frac{V'}{V}\right)^2 \gg \frac{T(y)}{V} \,,
\end{eqnarray}
so that for a fixed potential, DBI inflation could be achieved by appropriately reducing the background warp factor $T(y)$. However, microphysical constraints prevent the warp factor from becoming arbitrarily small: when the infrared scale of a throat becomes small compared to the scale of supersymmetry breaking (due to e.g.\ fluxes or antibranes in a different region of the compactification), then relevant supersymmetry-breaking perturbations of the throat sourced in the ultraviolet lead to large corrections to the infrared geometry (cf.\ the discussion in \cite{Baumann:2010sx}). This constraint enforces
\begin{eqnarray}
V \lesssim 2T_{3} \,a_{0}^{4} \equiv 2 T(y)|_{\rm tip} \le 2 T(y) \,.
\end{eqnarray}
For comparison, the general arguments of \cite{Baumann:2008kq} concerning the scale of compactification corrections give $V_{\rm bulk} \sim T_{3} a_{0}^{4}$, and our choice to expand around $Q = 1$ is consistent with these results.
We have taken $T_3=10^{-2}$ and $a_{0} = 10^{-3}$ throughout, so the condition (\ref{steep}) could only be satisfied if a rather steep potential arose by chance.
In fact, an additional effect reduces the likelihood of DBI inflation in our analysis. For consistency we have restricted our numerical evolution to the region where the singular conifold approximation is applicable, which excludes the region of greatest warping, the tip of the deformed conifold. In practice, we impose $x > 20\,a_{0}$, so that the minimum value of $T(y)$ explored in our simulations exceeds the global minimum value by a factor of $20^{4}$. It would be very interesting to extend our analysis to the tip region, and as the unperturbed Klebanov-Strassler solution is well understood, it would be straightforward to incorporate purely kinematic corrections involving deviations of the field space metric from that of the singular conifold. However, characterizing the structure of the potential in this region would be much more challenging, and would require understanding
the most general non-supersymmetric, perturbed solution for the tip region, along the lines of \cite{Baumann:2010sx} but departing from the approximately-conformal region.
One could also ask whether, even when the potential is not steep enough to accelerate the D3-brane to near the local speed of light, large initial velocities might still trigger a phase of DBI inflation.
We have found that cases with initial radial kinetic energy larger than 10\% of the initial potential energy typically strike the bottom without entering the DBI regime.
This result is compatible with prior investigations such as \cite{Bean:2007hc}, which found that exceptionally steep Coulomb potentials, which could arise in cones whose base spaces have extremely small angular volume, are required to produce DBI phases.
In summary, we find that the combination of the mild Coulomb potential in the Klebanov-Strassler throat, and general contributions to the D3-brane potential from moduli stabilization in the bulk, do not suffice to support DBI inflation in the $AdS_{5}\times T^{1,1}$ region. It would be interesting to understand whether DBI inflation arises in the tip region \cite{Kecskemeti:2006cg,Pajer:2008uy}.
\section{Towards the Primordial Perturbations}\label{experiment}
D3-brane inflation generically involves at least six\footnote{In addition to the six D3-brane coordinates, the compactification moduli can also evolve.} dynamical fields, and the study of the primordial perturbations is rather intricate. Most notably, entropy perturbations can be converted outside the horizon into curvature perturbations \cite{Gordon:2000hv}, provided that the inflaton trajectory bends in a suitable way.
In \S\ref{angularkinetic} we review the prospects for isocurvature-curvature conversion in warped D-brane inflation, following \cite{Chen:2008ada}. We will find that although our setup provides an efficient framework for computing multifield effects, there is a wide range of parameter space in which these effects can be neglected. Therefore, in \S\ref{effectivelysinglefield} we restrict our attention to the subset of cases that admit a single-field description, and straightforwardly obtain the CMB observables. A complete analysis of perturbations in the general case is postponed to a future publication.
\subsection{Angular kinetic energy and bending trajectories}\label{angularkinetic}
Typical trajectories that lead to prolonged inflation begin with relatively rapid angular and radial motion, then gradually spiral down to slow roll inflation along an inflection point, which is not necessarily parallel to the radial direction. Eventually, the D3-brane leaves the inflection point and accelerates, ending slow roll; it then plummets towards the anti-D3-brane, triggering tachyon condensation and annihilation of the brane-antibrane pair.
We will begin by assessing the prevalence of multifield effects in our ensemble of realizations.
As a first, rough measure of the importance of multiple fields, one can examine the angular kinetic energy during inflation.
In Figure \ref{fig:KE}, we show the ratio of angular kinetic energy to radial kinetic energy as a function of
the
number of e-folds before the end of inflation for selected examples.
In realizations yielding $N_{e}\approx 60$ e-folds, the ratio of angular kinetic energy to radial kinetic energy is often of order unity when observable scales exit the horizon, but diminishes thereafter. In realizations yielding $N_{e}\gtrsim 120$ e-folds, the transients are much diminished, but we still find cases (cf.\ Figure \ref{fig:KE}) in which the angular kinetic energy is of order the radial kinetic energy, and is approximately constant, throughout inflation. These cases involve slow roll inflation along an inflection point that is not parallel to the radial direction.
Although some degree of bending is commonplace, we do not find that the inflaton trajectory is substantially lengthened as a result of meandering \cite{Tye:2009ff} in six dimensions.
We find that the total distance $\ell$ in field space traversed in a realization with six active fields is negligibly larger than that for a realization with only one active field, $\ell(N_{f}=6) \lesssim 1.01 \ell(N_{f}=1)$.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.16in,angle=0]{KE.pdf}
\includegraphics[width=3.0in,angle=0]{KE_histogram_60_120.pdf}
\caption{The ratio ${\rm{KE}}_{\Psi}/{\rm{KE}}_{r}$ of angular to radial kinetic energies for trials with $\Delta_{\rm max} = 7.8$ and $N_{f} = 6$.
[Left panel] Evolution of ${\rm{KE}}_{\Psi}/{\rm{KE}}_{r}$ for trials yielding $60 \le N_e \lesssim 120$ e-folds of inflation (red lines), and $120 \le N_{e} \lesssim 180$ e-folds (black lines).
Notice that in some cases the angular kinetic energy is non-negligible, and nearly constant, in the final 60 e-folds, corresponding to an inflection point trajectory that is not purely radial.
[Right panel] Histogram of ${\rm{KE}}_{\Psi}/{\rm{KE}}_{r}$ 60 e-folds before the end of inflation for potentials that yield more than 60 (light blue) or more than 120 (dark blue) e-folds of inflation.}
\label{fig:KE}
\end{center}
\end{figure}
Next, we turn to a more precise characterization of multifield contributions to the primordial perturbations. A comprehensive study of multifield effects in D-brane inflation \cite{Chen:2008ada} has been performed in the framework of \cite{Baumann:2007ah}, i.e.\ in terms of explicit embeddings of D7-branes in the Klebanov-Strassler solution. One important lesson of \cite{Chen:2008ada} concerns the necessary conditions for isocurvature-curvature conversion at the end of inflation to make a significant contribution to CMB temperature anisotropies. Under fairly general assumptions, this contribution is negligible unless slow roll persists into the deformed conifold region, and the Coulomb potential is subdominant to the moduli potential at the time of tachyon condensation \cite{Chen:2008ada}.
Our analysis applies only in the region above the tip of the deformed conifold, so we cannot consistently capture large multifield effects from the {\it{end}} of inflation.\footnote{As explained in \S\ref{dbi}, incorporating these effects would require an extension of the results of \cite{Baumann:2010sx} to the tip region, which is beyond the scope of this work.}
A further possibility is that a sharp bend in the trajectory partway through inflation will produce substantial isocurvature-curvature conversion and render invalid a single-field treatment of the perturbations.
To quantify the contributions from additional fields, we calculate $\eta$ in two components, $\eta^\parallel$ and $\eta^\perp$, as defined in \cite{GrootNibbelink:2001qt} and \cite{vanTent:2001sk}.
The acceleration of the inflaton parallel to its instantaneous trajectory is captured by $\eta^\parallel$, while $\eta^\perp$ encodes the rate at which the inflaton trajectory bends perpendicular to itself. Therefore, $\eta^\perp$ is an efficient measure of the role of multiple fields in producing the primordial perturbations \cite{GrootNibbelink:2001qt}.
We define as ``effectively single-field" a realization in which the stringent cut $\eta^\perp/\eta^\parallel < 0.05$ is obeyed for the entirety of the last 60 e-folds.
Interestingly, we do find that in a small fraction of cases, abrupt angular motion occurs after a period of inflation driven by radial motion: the inflaton shifts rapidly from one angular minimum to another, then resumes radially-directed inflation. These examples with $\eta^\perp/\eta^\parallel \gg 0.05$ require a full multifield treatment of the perturbations, and there is the intriguing possibility of substantial non-Gaussianity from superhorizon evolution of isocurvature perturbations. We defer consideration of these interesting cases to a dedicated analysis \cite{forthcoming}.
\subsection{Single-field treatment of the perturbations}\label{effectivelysinglefield}
We now consider observational constraints on the substantial fraction of examples in which $\eta^\perp/\eta^\parallel < 0.05$, so that the primordial perturbations are well-approximated by the single-field result.
We begin by computing the Hubble slow roll parameters,
\begin{equation}
\epsilon \equiv -\frac{\dot{H}}{H^2}=-\frac{d\log H}{d N_e}\,,
\end{equation}
\begin{equation}
\eta \equiv \epsilon-\frac{1}{2}\frac{d\log \epsilon}{d N_e}\,,
\end{equation}
We describe the power spectrum of curvature fluctuations using a normalization $A_s$ and scalar spectral index, or `tilt', $n_s$, $\Delta_{{\mathcal R}}^{2}(k) = A_s (k/k_0)^{n_s-1}$.
In terms of $\eta_{60}$ and $\epsilon_{60}$, one has $A_s=\frac{V_{60}}{24\pi^2\epsilon_{60}}$ and $n_s-1=2\eta_{60}-4\epsilon_{60}$, where the subscripts denote evaluation 60 e-folds before the end of
inflation.
The scalar power has significant dependence on the parameter $D_{0}=2a_0^4 T_3$, which measures the height of the Coulomb potential. We therefore scan over a range of values of $D_0$ (in practice, we fix $T_3$ and scan over $a_{0}$), and for each successful trial that yields at least 60 e-folds, we compute the slow roll parameters, $A_{s}$, and $n_{s}$.
In Figure \ref{fig:epsandeta}, we show scatter plots of $\epsilon_{60}$ and $\eta_{60}$ as a function of the maximum number of e-folds.\footnote{Figures \ref{fig:epsandeta}, \ref{fig:power} and \ref{fig:rns} share a set of $4.9 \cdot 10^{6}$ Monte Carlo trials at $\Delta_{\rm max}=6$,
out of which 8301 trials yield more than 60 e-folds and 140 also satisfy the WMAP7 constraints on $A_{s}$ at $2\sigma$. Figure \ref{fig:rns} additionally includes $5 \cdot 10^{5}$ trials at $\Delta_{\rm max}=7.8$, out of which 750 examples yield more than 60 e-folds and 9 examples also satisfy the constraints on $A_{s}$ at $2\sigma$. All the data points given in Figures \ref{fig:epsandeta} and \ref{fig:power} obey $\eta^\perp/\eta^\parallel < 0.05$, while in
Figure \ref{fig:rns}, data points with $\eta^\perp/\eta^\parallel \ge 0.05$ are included, and indicated by red or purple dots.}
Notice the paucity of examples with $N_{e} \lesssim 120$. As $P(N_e) \propto N_e^{-3}$, a large fraction of trials yield $N_{e}$ in this range, but most such examples are excluded by the cut $\eta^\perp/\eta^\parallel < 0.05$. This can be understood from Figure \ref{fig:KE}: realizations yielding $N_{e} \ll 120$ typically have substantial angular evolution in the final 60 e-folds, so that the single-field approximation is inapplicable.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.0in,angle=0]{eps11.png}
\includegraphics[width=3.0in,angle=0]{eta11.png}
\caption{Hubble slow roll parameters $\epsilon_{60}$ and $\eta_{60}$ as a functions of the maximum number of e-folds. The scatter in $\epsilon_{60}$ for $N_{e} \gtrsim 120$ results from our scan over different values of the inflationary scale, as encoded in $a_0$. Notice the absence of corresponding scatter in $\eta_{60}$. Color coding: $-5 < \log_{10} a_0 < -4.75$, green; $-4.75 < \log_{10} a_0 < -4.5$, chartreuse; $-4.5 < \log_{10} a_0 < -4.25$, orange. All points shown have trajectories with negligible bending, $\eta_{\perp} < .05 \eta_{\|}$.}
\label{fig:epsandeta}
\end{center}
\end{figure}
We observe that $\eta_{60}$ is strongly correlated with the number of e-folds.
Importantly, when the single-field slow roll approximation is valid, only cases with $N_{e} \gtrsim 120$ e-folds are observationally consistent, since it is only for these cases that we have $V^{\prime\prime}<0$, ensuring $n_s<1$. This is not surprising (cf.\ \cite{Baumann:2007ah}, Figure \ref{fig:inflection}): for single-field inflation in an approximate inflection point that is flat enough to yield exactly 60 e-folds of inflation, the CMB anisotropies are generated when the inflaton is above the inflection point, so that the potential is concave up, and hence $n_{s}>1$.
In a corresponding potential that yields 120 e-folds of inflation, observable modes exit the horizon when the inflaton is near to the inflection point (because 60 e-folds have elapsed and 60 e-folds remain), so that for $\epsilon_{60} \ll 1$, one has $n_{s} \approx 1$.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.5in,angle=0]{inflection.png}
\caption{An inflection-point potential in one dimension, taken from \cite{Baumann:2007ah}. The color coding indicates that if the inflaton is above the inflection point 60 e-folds before the end of inflation, $V^{\prime\prime}>0$ and the scalar power spectrum is blue. A red spectrum is possible if the inflaton has passed the inflection point 60 e-folds before the end of inflation. }
\label{fig:inflection}
\end{center}
\end{figure}
The seven-year WMAP (WMAP7) constraints on $A_{s}$ and $n_{s}$ are $A_{s}=(2.43\pm0.11)\times 10^{-9}$ and $n_s=0.963\pm0.014$ at
$k=0.002$ Mpc$^{-1}$, at the 68\% confidence level \cite{Larson:2010gs}.
In Figure \ref{fig:power} we show a scatter plot of $A_s$ compared to the measured central value, $A_{s}^{\star}=2.43\times 10^{-9}$, indicating cases that are allowed at $2\sigma$ by the
WMAP7 constraint.
Notice that there is no fine-tuned choice of $D_0$, or of other input parameters, that guarantees a WMAP-normalized spectrum of perturbations: there is significant dispersion in $\epsilon_{60}$ in
our ensemble of inflationary models, with corresponding dispersion in the scalar power.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=3.0in,angle=0]{power11.png}
\includegraphics[width=3.0in,angle=0]{powerzoomin11.png}
\caption{Scalar power $A_s$ compared to the measured central value, $A_{s}^{\star}=2.43\times 10^{-9}$, as a function of the total number of e-folds. On the right, we show a zoomed in version. Within the red lines, the scalar power is consistent with WMAP7 at $2\sigma$. All points shown have $\eta_{\perp} < .05 \eta_{\|}$, and the color coding is as in Figure \ref{fig:epsandeta}.}
\label{fig:power}
\end{center}
\end{figure}
Next, for the subset of cases in which $A_{s}$ is consistent with experiment, we calculate the corresponding tilt $n_s$ and tensor-to-scalar ratio $r$, displaying the results in Figure \ref{fig:rns}.
Evidently, observational constraints on the tilt are readily satisfied for essentially all cases with $N_{e} \gtrsim 120$ e-folds, while cases with $60 \lesssim N_{e} \lesssim 120$ e-folds solve the horizon and flatness problems but are not consistent with experiment. We stress that this is a straightforward, albeit interesting, consequence of the inflection point form of the potential that is characteristic of successful realizations in our ensemble.
Within our model there is a microphysical upper bound $r \le r_{\rm max}$ on the tensor-to-scalar ratio resulting from the geometric bound on $\phi_{\rm UV}$ \cite{Baumann:2006cd}. However,
in single-field slow roll cases whose scalar perturbations are consistent with WMAP7, we find that $r \ll r_{\rm max}$.
It is sometimes argued that small values of $r$ require substantial fine-tuning. Evaluating the absolute likelihood of inflation in this scenario requires information about a priori measures, and is beyond the scope of this work. Even so, our analysis indicates that the requisite degree of fine-tuning is not extreme: in optimal regions of the parameter space, but without direct fine-tuning of the potential, we find examples consistent with all observations approximately once in $10^{5}$ trials, and it is clear from Figure \ref{fig:rns} that these scenarios have $r \lesssim 10^{-12}$.
\begin{figure}[!h]
\begin{center}
\includegraphics[width=4.5in,angle=0]{rns11.png}
\caption{Scalar tilt $n_s$ and tensor-to-scalar ratio $r$ for cases for which the scalar power $A_s$ is consistent with WMAP7 at $2\sigma$. The red lines indicate the region allowed at $2\sigma$ by the WMAP7 constraints on $n_s$. Green and red dots arise from realizations with $\Delta_{\rm max}=6$, while cyan and purple dots have $\Delta_{\rm max}=7.8$. Green and cyan dots correspond to trajectories with negligible bending ($\eta_{\perp} < .05 \eta_{\|}$), while red and purple dots have $\eta_{\perp} \ge .05 \eta_{\|}$ and plausibly require a multifield analysis. }
\label{fig:rns}
\end{center}
\end{figure}
\section{Conclusions} \label{conclusions}
We have performed a comprehensive Monte Carlo analysis of D3-brane inflation for a general scalar potential on the conifold, obtaining robust predictions in spite of --- indeed, arguably because of --- the complexity of the inflaton action. Our work builds on recent results \cite{Baumann:2010sx} that provide the structure of the potential, i.e.\ a list of possible terms in the potential with undetermined coefficients. Most previous works (cf.\ e.g. \cite{Burgess:2006cb,Bean:2007hc,Baumann:2007np,Krause:2007jk,Baumann:2007ah,Ali:2010jx}) have treated special configurations in which suitably-aligned D7-branes stabilize the angular positions of the inflationary D3-brane, leading to one-field or two-field dynamics. We have characterized the six-dimensional dynamics of the homogeneous background, without restricting the scalar potential. We have not fine-tuned the potential by hand, but have instead drawn the coefficients in the scalar potential from suitable distributions, creating an ensemble of potentials, a subset of which led to inflation by chance.
We found that the probability $P(N_e)$ of $N_e$ e-folds of inflation is a power law, $P(N_{e}) \propto N_{e}^{-\alpha}$, with $\alpha \approx 3$. The exponent is robust against changes in our truncation of the potential, in the statistical distribution from which the coefficients in the potential are drawn, in the initial conditions, and in the number of dynamical fields.
Moreover, we derived $\alpha=3$ from a simple analytical model of inflection point inflation. This power-law behavior has significant implications for the prospect of detecting transients arising from the onset of inflation (cf.\ \cite{Freivogel:2005vv}): among all histories with at least 60 e-folds, histories with at least 65 e-folds --- in which most transients are stretched to unobservable scales --- are considerably more likely.
When the inflaton starts at a radial location that is far above the inflection point, angular motion combined with gradual radial infall frequently allow the inflaton to reach the inflection point with a velocity small enough to permit inflation.
Moreover, we found attractor behavior in the angular directions: in an order-unity fraction of the space of initial angular positions, the inflaton spirals down to the inflection point. However, large amounts of radial or angular kinetic energy, of order the initial potential energy, are compatible with inflation only in exceptional cases.
DBI inflation did not arise by chance in our ensemble: the potential was never steep enough. It would be interesting to understand whether this finding can be generalized or is an artifact of the limitations of our treatment.
We have obtained the scalar perturbations for the subset of realizations in which a single-field description is applicable throughout the final 60 e-folds, deferring a comprehensive study of the multifield evolution of perturbations to future work.
In optimal regions of the parameter space, we found that 60 or more e-folds of inflation arose approximately once in $10^3$ trials, but because constraints on $A_s$ and $n_s$ enforce $N_e \gtrsim 120$, observational constraints were satisfied approximately once in $10^5$ trials.
Outside the optimal regions, the chance of inflation diminished rapidly. As we lack a meaningful {\it{a priori}} measure on the space of parameters and initial conditions, we have not attempted to quantify the total degree of fine tuning, but our results provide considerable information about relative likelihoods.
In the range of parameters where realizations consistent with observations of the scalar power spectrum are most likely, we found that the tensor-to-scalar ratio obeys $r \lesssim 10^{-12}$
in all examples allowed by WMAP7, which is much smaller than the maximum allowed by the Lyth bound.
Our statements about the perturbations apply only to realizations that are consistently described by slow roll, effectively single-field inflation, which we checked by computing the rate of bending of the trajectory. We anticipate that including more general multifield cases could populate additional regions of the $n_{s}-r$ plane.
Our findings have interesting implications beyond the setting of D-brane inflation. The fact that our conclusions are unaffected by the statistical distribution used to generate the coefficients in the potential suggests that in inflationary models in which there are many competing terms in the scalar potential, the details of the individual terms
can be less important than the collective structure. A simplification of this form has been previously noted \cite{RMTNflation} in inflation driven by $D \gg 1$ fields with quadratic potentials \cite{Nflation}, where random matrix theory could be applied directly.{\footnote{See \cite{Frazer:2011tg} for recent related work.}} Our present results suggest that this emergent simplicity may be more general, and it would be valuable to understand whether a general inflationary model in a field space of dimension $D \gg 1$ has characteristic properties at large $D$.
\subsection*{Acknowledgements}
We are grateful to Thomas Bachlechner, Daniel Baumann, Richard Easther, Raphael Flauger, Ben Freivogel, Sohang Gandhi, Ben Heidenreich, Enrico Pajer, Gary Shiu, Henry Tye, Timm Wrase, and Jiajun Xu for helpful discussions. We particularly thank Dan Wohns for extensive discussions and suggestions, and Daniel Baumann and Raphael Flauger for their useful comments on the manuscript.
N.A. and R.B.'s research is supported by NSF CAREER grant AST0844825, NSF grant PHY0555216, NASA Astrophysics Theory Program grant NNX08AH27G and by Research Corporation. The research of L.M. and G.X. was supported by the Alfred P. Sloan Foundation and by the NSF under grant PHY-0757868. G.X. was supported in part by the Institute for Advanced Study, Hong Kong University of Science and Technology. L.M. thanks the organizers of the Primordial Features and Non-Gaussianities workshop at HRI for providing a stimulating environment during the course of this work.
|
\section{Introduction}
Active research on graphene\cite{novo1} revealed not only
numerous exceptional properties
\cite{geim,berger,kats,hasan2010} but also have prepared a
ground for the discovery of several graphene based materials.
Preparation of freestanding graphene sheets with nonuniform
oxygen coverage have been achieved.\cite{dikin} More
recently the synthesis of two-dimensional hydrocarbon in
honeycomb structure, so called \textit{graphane}
\cite{novo-graphane} (CH), showing diverse electronic, magnetic
and mechanical properties
\cite{sofo,boukhvalov,graphane-h1,graphane-h2,graphane-m} is
reported.
According to Pauling scale F has electronegativity of 3.98, which is
higher than those of C(2.55), H(2.20) and O(3.44), and hence
fluorination of graphene is expected to result in a material,
which may be even more interesting than both graphene oxide
and CH. Before the first synthesis of graphene, fluorinated
graphite has been treated theoretically.
\cite{charlier,takagi} Owing to promising properties
revealed for CH, fluorinated graphene structures are now
attracting considerable interest
\cite{boukhvalov2,cheng,robinson,nair,yeni1, yeni2, yeni3,
yeni4} despite uncertainties in their chemical compositions
and atomic structures. In an effort to identify the
structures of fluorinated samples, previous theoretical
models attempted to deduce the lowest energy structures.
\cite{charlier,boukhvalov2} In addition, band gaps of
different structures calculated within Density Functional
Theory (DFT) are compared with the values revealed
through specific measurements.\cite{robinson,nair} However,
neither the stability of proposed structures are questioned,
nor underestimation of band gaps within DFT has been a
subject matter. Raman spectrum by itself, has been limited in
specifying C$_n$F structures.\cite{nair}
In this work, we first determined stable C$_n$F structures
for $n \leq 4$. Then we revealed specific properties (such as
internal structural parameters, elastic constants, formation
and binding energies, energy band gap and photoelectric
threshold) for those stable structures as signatures to
identify the derivatives probed experimentally. We placed an
emphasis on fully fluorinated graphene or fluorographene
(CF), in which D and G Raman peaks of bare graphene disappear
after long fluorination period.\cite{robinson,nair} Present
study reveals that the properties, such as structural
parameters, binding energy, band gap and phonon modes of
various fluorinated structures are strongly dependent on the
binding structure of F atoms and their composition. Some of
these properties are found to roughly scale with F coverage.
While the stable C$_2$F chair structure is metallic, CF is a
nonmagnetic insulator with a band gap, $E_{g}$, being much
larger than 3 eV, i.e. a value attributed experimentally to
fully fluorinated graphene. In view of the calculated
diffusion constant, Raman active modes and other properties,
available experimental data suggest that domains (or grains)
of various C$_n$F structures with extended and imperfect
grain boundaries can coexist after fluorination process.
Hence the measured properties are averaged from diverse
perfect and imperfect regions.
\section{Computational Methodology}
Our predictions are obtained from first-principles plane wave
calculations\cite{vasp} within DFT, which is demonstrated to
yield rather accurate results for carbon based materials.
Calculations are performed using spin-polarized local density
approximation (LDA)\cite{lda} and projector augmented wave
(PAW) potentials.\cite{paw} Kinetic energy cutoff $ \hbar^2 |
\mathbf{k}+\mathbf{G}|^2 / 2m $ for plane-wave basis set is
taken as 500 eV. In the self-consistent potential and total
energy calculations of fluorographene a set of (25x25x1)
\textbf{k}-point sampling is used for Brillouin zone (BZ)
integration. The convergence criterion of self consistent
calculations for ionic relaxations is $10^{-5}$ eV between
two consecutive steps. By using the conjugate gradient
method, all atomic positions and unit cells are optimized
until the atomic forces are less than 0.03 eV/\AA. Pressures
on the lattice unit cell are decreased to values less than
0.5 kBar. The energy band gap, which is usually
underestimated in DFT, is corrected by frequency-dependent
GW$_{0}$ calculations.\cite{gw} In GW$_{0}$ corrections
screened Coulomb potential, W, is kept fixed to initial DFT
value W$_{0}$ and Green's function, G, is iterated four
times. Various tests are performed regarding vacuum spacing,
kinetic energy cut-off energy, number of bands, \textbf{k}-points
and grid points. Finally, the band gap of CF is found
7.49 eV after GW$_{0}$ correction, which is carried out by
using (12x12x1) \textbf{k}-points in BZ, $15$~\AA~ vacuum
spacing, default cut-off potential for GW$_{0}$, 192 bands
and 64 grid points. Phonon frequencies and phonon
eigenvectors are calculated using the Density Functional
Perturbation Theory (DFPT).\cite{pwscf}
\section{Structures of fluorinated graphene}
Each carbon atom of graphene can bind only one F atom and
through coverage (or decoration) of one or two sides of
graphene, one can achieve diverse C$_{n}$F structures.
Uniform F coverage is specified by $\Theta = 1/n$ (namely one
F adatom per $n$ C atoms), whereby $\Theta = 0.5$ corresponds
to half fluorination and $\Theta = 1$ is fluorographene CF.
The adsorption of a single F atom to graphene is precursor
for fluorination. When placed at diverse sites of a $(4
\times 4)$ supercell of graphene, simple F atom moves to the
top site of a carbon atom and remains adsorbed there. For the
resulting structure spin-polarized state (with 0.4 Bohr magneton) is only 2 meV favorable than the nonmagnetic state. The binding energy of F is
$E_b$=2.71 eV in equilibrium, that is a rather strong binding
unlike many other adatoms adsorbed to graphene. An energy
barrier, $Q_{B} =\sim$0.45 eV, occurs along its minimum
energy migration path. Our calculations related with the
minimum energy path of a single F atom follow hexagons of
underlying graphene. Namely, F atom migrates from the highest
binding energy site, i.e. top site (on top of carbon atom)
to the next top site through bridge site (bridge position
between two adjacent carbon atoms of graphene). The
corresponding diffusion constant for a single F atom, $D= \nu
a e^{-Q_{B}/k_{B}T}$ is calculated in terms of the lattice
constant, $a=2.55$ \AA~and characteristic jump frequency
$\nu\approx$39 THz. Experiments present evidences
that energy barriers on the order of 0.5 eV would make the
adatoms mobile.\cite{nair,ref1} Moreover, this energy barrier is
further lowered even it is collapsed in the presence of a
second F atom at the close proximity. Consequently, this
situation together with the tendency towards clustering
favors that C$_n$F grains (or domains) of different $n$
on graphene can form in the course of fluorination. We note
that the energy barrier for the diffusion of a single carbon
adatom adsorbed on the bridge sites of graphene was
calculated to be in the similar energy range. Carbon adatoms on
graphene were found to be rather mobile. That energy barrier
for single C adatom was found to decrease, even to collapse at
the close proximity of a second adatom.\cite{carbon-adatom}
\begin{figure}
\centering
\includegraphics[width=8.5cm]{f1.png}
\caption{(Color online) Atomic structure and calculated
phonon bands (i.e. phonon frequencies versus wave vector, \textbf{k}) of various optimized C$_n$F
structures calculated along the symmetry directions of BZ.
Carbon and fluorine atoms are indicated by black (dark) and
blue (light) balls, respectively. (a) C$_2$F Chair structure.
(b) C$_2$F Boat structure. (c) C$_4$F structure. Units are
\AA~for structural parameters and cm$^{-1}$ for frequencies.}
\label{fig1}
\end{figure}
\begin{table*}
\caption{Comparison of the calculated properties of four
stable, fluorinated graphene structures (namely CF, C$_2$F
chair, C$_2$F boat and C$_4$F) with those of graphene and CH.
Lattice constant, $a=b$ ($a \neq b$ for rectangular lattice);
C-C bond distance, $d_{CC}$ (second entries with slash differ
from the previous one); C-X bond distance (X indicating H (F)
atom for CH (CF)), $d_{CX}$; the buckling, $\delta$; angle
between adjacent C-C bonds, $\alpha_{C}$; angle between
adjacent C-X and C-C bonds, $\alpha_{X}$; total energy per
cell comprising 8 carbon atoms $E_T$; formation energy per X
atom relative to graphene, $E_f$; binding energy per X atom
relative to graphene, $E_b$ (the value in parenthesis
$E_{b^{'}}$ excludes the X-X coupling); desorption energy,
$E_d$ (see the text for formal definitions); energy band gap
calculated by LDA, $E_{g}^{LDA}$; energy band gap corrected
by GW$_{0}$, $E_{g}^{GW_{0}}$; photoelectric threshold,
$\Phi$; in-plane stiffness, $C$; Poison ratio, $\nu$. All
materials are treated in hexagonal lattice, except C$_2$F
boat having rectangular lattice.}
\label{table} \centering{}\begin{tabular}{lcccccccccccccccc}
\hline \hline
Material & $a$ $(b)$ & $d_{CC}$ & $d_{CX}$ & $\delta$ &
$\alpha_{C}$ & $\alpha_{X}$ & $E_{g}^{LDA}$ & $E_{g}^{GW_{0}}$ &
$E_{T}$ & $E_{f}$ &$E_{b}$$(E_{b^{'}})$& $E_{d}$& $\Phi$& $C$ &
$\nu$ \tabularnewline
& (\AA{}) &(\AA{}) & (\AA{}) & (\AA{}) &
(deg) & (deg) & ($eV$) &($eV$) & ( $eV$ ) & ($eV$) & ($eV$)
& ($eV$) &($eV$)&($J/m^{2}$) & \tabularnewline \hline
Graphene\cite{h-ansiklopedi} & 2.46 &1.42 & - & 0.00 & 120
& - & 0.00 & 0.00 &-80.73 & - & - & - & 4.77 & 335 & 0.16
\tabularnewline \hline
CH\cite{graphane-h1} &2.51 & 1.52 & 1.12 & 0.45 & 112 & 107 &
3.42 & 5.97 & -110.56 & 0.39 & 2.8(2.5) & 4.8 &4.97 & 243 & 0.07
\tabularnewline \hline
CF & 2.55 & 1.55 & 1.37 & 0.49 & 111 & 108 & 2.96 &
7.49 & -113.32 & 2.04 & 3.6(2.9) & 5.3 &7.94 & 250 & 0.14
\tabularnewline \hline
C$_2$F chair & 2.52 & 1.48 & 1.47 & 0.29 & 116 & 101 & metal
& metal & -89.22 & 0.09& 1.7(0.9) & 1.2 &8.6/5.6 & 280 & 0.18
\tabularnewline \hline
C$_2$F boat & 2.54(4.36) & 1.51/1.61 & 1.40 & 0.42 & 114/118 &
100/101 & 1.57 & 5.68 & -92.48 & 0.91 & 2.5(1.6) & 2.4 &7.9/5.1 &
286(268) & 0.05 \tabularnewline \hline
C$_{4}$F & 4.92 & 1.49/1.39 & 1.43 & 0.34 & 114/119 & 104
& 2.93 & 5.99 & -87.68 &1.44 & 3.0(2.7)& 3.5 & 8.1/5.6 & 298 & 0.12
\tabularnewline \hline \hline
\end{tabular}
\end{table*}
In earlier theoretical studies,
\cite{charlier,boukhvalov2,robinson}
the total energies and/or binding energies were taken as
criteria for whether a given C$_n$F structure exists. Even if
a C$_n$F structure seems to be in a minimum on the Born-
Oppenheimer surface, its stability is meticulously examined
by calculating frequencies of all phonon modes in BZ. Here we
calculated phonon dispersions of most of optimized C$_n$F
structures. We found C$_4$F, C$_2$F boat, C$_2$F chair (See
Fig.~\ref{fig1}) and CF chair (See Fig.~\ref{fig2})
structures have positive frequencies throughout the Brillouin
zone indicating their stability.
Some of phonon branches of C$_n$F structures (for example, CF
boat) have imaginary frequencies and hence are unstable, in
spite of the fact that their structures can be optimized. The
possibility that these unstable structures can occur at
finite and small sizes is, however, not excluded. For stable
structures, the gap between optical and acoustical branches
is collapsed, since the optical branches associated with the
modes of C-F bonds occur at lower frequencies. This situation
is in contrast with the phonon spectrum of graphane,
\cite{graphane-h1} where optical modes related with C-H bonds
appear above the acoustical branches at $\sim$ 2900
cm$^{-1}$.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{f2.png}
\caption{(Color online) (a) Atomic structure of
fluorographene CF. $a$ and $b$ are the lattice vectors ($|a|
=|b|$) of hexagonal structure; $d_{CC}$ ($d_{CF}$) is the C-C
(C-F) bond distance; $\delta$ is the buckling. (b) Phonon
frequencies versus wave vector \textbf{k} of optimized CF calculated along
symmetry directions in BZ. (c) Symmetries, frequencies and
descriptions of Raman active modes of CF. (d) Calculated
Raman active modes of graphene, CH, CF and C$_4$F are
indicated on the frequency axis. Those modes indicated
by "+" are observed experimentally. There is no experimental
Raman data in the shaded regions. Units are \AA~for
structural parameters and cm$^{-1}$ for frequencies.}
\label{fig2}
\end{figure}
The formation energy of fluorination is defined as
$E_f=(n_{F_{2}} E_{T,F_{2}}+E_{T,Gr}-E_{T,C_{n}F})/n_F$ in
terms of the total ground state energies of optimized
structures of graphene and fluorinated graphenes at different
composition, respectively, $E_{T,Gr}$, $E_{T,C_{n}F}$, and
the total ground state energy of single carbon atom,
$E_{T,C}$, of F$_2$ molecule and F atom, $E_{T,F_{2}}$ and
$E_{T,F}$. Similarly, the binding energy of F atom relative
to graphene including F-F coupling is $E_{b}=(E_{T,Gr}+n_{F}
E_{T,F}- E_{T,C_{n}F})/n_{F}$ and without F-F coupling
$E_{b^{'}}=(E_{T,Gr}+E_{T,n_{F}F}- E_{T,C_{n}F})/n_{F}$. Here
$E_{T,n_{F}F}$ is the total energy of suspended single or
double layers of F occupying the same positions as in C$_n$F.
The desorption energy, $E_d$ is the energy required to remove
one single F atom from the surface of C$_n$F. $n_{F_{2}}$ and
$n_F$ are numbers of F$_2$ molecules and F atoms,
respectively. The total energies are calculated in
periodically repeating supercells comprising 8 carbon atoms
and keeping all the parameters of calculations described
above using spin-polarized as well as spin unpolarized LDA.
Lowest (magnetic or nonmagnetic) total energy is used as
ground state total energy.
Fluorographene (CF), where F atoms are bound to each C atom
of graphene alternatingly from top and bottom sides is
energetically most favorable structure. Upon full
fluorination, the planar honeycomb structure of C atoms
becomes buckled (puckered) and C-C bond length increases by
$\sim$10\%. At the end, while planar $sp^2$-bonding of
graphene is dehybridized, the buckled configuration is
maintained by $sp^3$-like rehybridization. In
Table~\ref{table}, the calculated lattice constants, internal
structural parameters, relevant binding energies and energy
band gaps of stable C$_n$F structures are compared with those
of bare graphene and CH.\cite{graphane-h1} Notably, internal
parameters (such as $\delta$, C-C bond length) as well as
lattice constants of various C$_n$F structures vary with F
coverage, $\Theta$. CF, has highest values for $E_f$, $E_b$,
$E_{b^{'}}$ and $E_d$ given in Table~\ref{table}; those of
C$_4$F are second highest among stable C$_n$F structures.
Since Raman spectrum can convey information for a particular
structure and hence can set its signature, calculated Raman
active modes of stable C$_4$F and CF structures together with
those of graphene and CH are also indicated in
Fig.~\ref{fig2} (c) and (d). It is known that the only
characteristic Raman active mode of graphene at 1594
cm$^{-1}$ is observed so far.
\cite{r-graphene} Similarly, for CH the mode at $\sim$1342
cm$^{-1}$ is observed.\cite{novo-graphane} One of two
Raman active modes of C$_4$F at 1645 cm$^{-1}$ seems to be
observed.\cite{robinson} In compliance with the theory,
\cite{ferrari} phonon branches of all these observed modes
exhibit a kink structure. However, none of the Raman active
modes of CF revealed in Fig.~\ref{fig2} has been observed
yet. Raman spectroscopy in the low frequency range may be
useful in identifying experimental structures.
\section{Electronic Structures}
Energy bands, which are calculated for the optimized
C$_4$F, C$_2$F boat, C$_2$F chair and CF chair structures are
presented in Fig.~\ref{fig3} and Fig.~\ref{fig4} structures,
respectively. The orbital projected densities of states (PDOS)
together with the total densities of states of these
optimized structures are also presented. Analysis of the
electronic structure can also provide data to reveal the
observed structure of fluorinated graphene. As seen in
Table~\ref{table}, stable C$_n$F structures have LDA band
gaps ranging from 0 eV to 2.96 eV. Surprisingly, C$_2$F chair
structure is found to be a metal owing to the odd number of
valance electrons in the primitive unit cell. Even if various
measurements on the band gap of fluorinated graphene lie in
the energy range from 68 meV\cite{cheng} to 3 eV,\cite{nair}
these calculated band gaps are underestimated by LDA.
Incidentally, the band gaps change significantly after they
are corrected by various self-energy methods. In fact, the
correction using GW$_0$ self-energy method predicts a rather
wide band gap of 7.49 eV for CF. The corrected band gaps for
C$_2$F boat structure and C$_4$F are 5.68 eV and 5.99 eV,
respectively. It should be noted that the GW$_0$ self-energy
method has been successful in predicting the band gaps of 3D
semiconductors.\cite{hse06}
\begin{figure}
\centering
\includegraphics[width=8.5cm]{f3.png}
\caption{(Color online) Energy band structures of various
stable C$_n$F structures together with the orbital projected
densities of states and the total densities of states (DOS).
The LDA band gaps are shaded and the zero of energy is set to
the Fermi level $E_{F}$. Total DOS is scaled to 45\%. Valence
and conduction band edges after GW$_{0}$ correction are
indicated by filled/red circles. (a) C$_2$F chair structure.
(b) C$_2$F Boat structure. (c) C$_4$F structure.}
\label{fig3}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=8.5cm]{f4.png}
\caption{(Color online) (a) Energy band structure of CF
together with the orbital projected DOS and
total densities of states. The LDA band gap is shaded and the
zero of energy is set to the Fermi level, $E_{F}$. Valence
and conduction band edges after GW$_{0}$ correction are
indicated by filled/red circles. (b) Isosurfaces of charge
densities of states corresponds to first (V1), second (V2)
valance and first (C1) and second (C2) conduction bands at
the $\Gamma$- and $K$-points. (c) Contour plots of the total
charge density $\rho_T$ and difference charge density
$\Delta\rho$ in the plane passing through F-C-C-F atoms.
Contour spacings are 0.03 e/\AA$^{3}$.}
\label{fig4}
\end{figure}
While predicting much larger band gap for CF, the measured
band gap of $\sim$3 eV reported by Nair et al.\cite{nair}
marks the serious discrepancy between theory and experiment.
The character of the band structure of CF are revealed from
the analysis of projected density of states as well as charge
densities of specific bands in Fig.~\ref{fig4} (b).
Conduction band edge consists of the antibonding combination
of $p_{z}$-orbitals of F and C atoms. The $p_{z}$-orbitals of
C atoms by themselves, are combined to form $\pi$-bands. The
bands at the edge of the valence band are derived from the
combination of C-$p_{x}+p_{y}$ and F-$p_{x}+p_{y}$ orbitals.
The total contribution of C orbitals to the valence band can
be viewed as the contribution of four tetrahedrally
coordinated $sp^{3}$-like hybrid orbitals of $s$- and
$p$-orbitals of C atoms. However, the deviation from
tetrahedral coordination increases when $n$ increases or
single side is fluorinated. As a matter of fact, the total
density of states presented in Fig.~\ref{fig3} and
Fig.~\ref{fig4} mark crucial differences. In this respect,
spectroscopy data is expected to yield significant
information regarding the observed structures of fluorinated
graphenes.
The contour plots of the total charge density, $\rho_T$, in
the F-C-C-F plane suggests the formation of strong covalent
C-C bonds from the bonding combination of two C-$sp^{3}$
hybrid orbitals. The difference charge density, $\Delta \rho$
(which is obtained by subtracting the charges of free C and
free F atoms situated at their respective positions in
CF), indicates charge transfer to the middle of C-C bond and
to F atom, revealing the bond charge between C atoms and
ionic character of C-F bond. However, the value of charge
transfer is not unique, but diversifies among different
methods of analysis.\cite{charge-analysis} Nevertheless, the
direction of calculated charge transfer is in compliance with
the Pauling ionicity scale and is corroborated by calculated
Born effective charges, which have in-plane ($\parallel$) and
out-of-plane ($\perp$) components on C atoms, $Z^{*}_{C,
\parallel}=0.30$, $Z^{*}_{C,\perp}=0.35$ and on F atoms
$Z^{*}_{F,\parallel}=-0.30$, $Z^{*}_{F,\perp}=-0.35$.
Finally, we note that perfect CF is a nonmagnetic insulator.
However, a single isolated F vacancy attains a net magnetic
moment of 1 Bohr magneton ($\mu_{B}$) and localized defect
states in the band gap. Creation of an unpaired $\pi$-
electron upon F vacancy is the source of magnetic moment.
However, the exchange interaction between two F-vacancies
calculated in a (7x7x1) supercell is found to be nonmagnetic
for the first nearest neighbor distances due to spin pairings.
Similar to graphane,\cite{graphane-h1,graphane-h2} it is also
possible to attain large magnetic moments on F-vacant domains
in CF structures.
\begin{figure*}
\centering
\includegraphics[width=14cm]{f5.png}
\caption{(Color online) (a) Variation of strain energy and
its first derivative with respect to the uniform strain
$\epsilon$. Orange/gray shaded region indicates the plastic
range. Two critical strains in the elastic range are labeled
as $\epsilon_{c1}$ and $\epsilon_{c2}$. (b) Variation of the
band gaps with $\epsilon$. LDA and $GW_0$ calculations are
carried out using 5x5 supercell having the lattice parameter
of c$_{0}$=5a, and $\Delta c$ is its stretching}
\label{figs5}
\end{figure*}
\section{Elastic Properties of CF}
Having analyzed the stability of various C$_n$F structures
with $n=$1,2 and 4, we next investigate their mechanical
properties. The elastic properties of this structure can
conveniently be characterized by its Young's modulus and
Poison's ratio. However, the in-plane stiffness $C$ is known
to be a better measure of the strength of single layer
honeycomb structures, since the thickness of the layer $h$
cannot be defined unambiguously. Defining $A_{0}$ as the
equilibrium area of a C$_n$F structures, the in-plane
stiffness is obtained as $C=(\delta E^{2}_{s}/ \delta
\epsilon^2)/A_{0}$, in terms of strain energy $E_s$
and uniaxial strain $\epsilon$.\cite{graphane-m} The values
of in-plane stiffness $C$, and Poisson's ratio $\nu$,
calculated for stable C$_n$F structures are given in
Table~\ref{table} together with the values calculated for
graphene and graphane. For example, the calculated values of
CF are $C$=250 J/m$^2$ and $\nu$=0.14. It is noted that $C$
increases with $n$. For CF (i.e. $n=1$), the in-plane
stiffness is close to that calculated for CH. It appears that
either the interaction between C-F bonds in CF (or the
interaction between C-H bonds in CH) does not have
significant contribution to the in-plane stiffness. The main
effect occurs through dehybridization of $sp^2$ bonds of
graphene through the formation C-F bonds (or C-H bonds).
A value of Young's modulus around 0.77 $TPa$ can be
calculated by estimating the thickness of CF as $h$=3.84
\AA{}, namely the sum of the thickness of graphene (3.35
\AA) and buckling, $\delta$ (0.49 \AA). This value is
smaller, but comparable with the value proposed for graphene,
i.e. $\sim$ 1 $TPa$. Here the contribution of C-F bonds to
the thickness of CF is neglected, since the interaction
between C-F bonds has only negligible effects on the strength
of CF.
In Fig.~\ref{figs5} the variation of strain energy $E_s$ and
its derivative, $\delta E_{s}/\delta \epsilon$, with strain,
$\epsilon$ are presented in both elastic and plastic
regions. Two critical strain values, $\epsilon_{c_{1}}$ and
$\epsilon_{c_{2}}$, are deduced. The first one,
$\epsilon_{c_{1}}$, is the point where the derivative curve
attains its maximum value. This means that the structure can
be expanded under smaller tension for higher values of
strain. This point also corresponds to phonon
instability\cite{graphane-m} where the longitudinal acoustic
modes start to become imaginary for $\epsilon >
\epsilon_{c_{1}}$. The second critical point,
$\epsilon_{c_{2}}$ $(\simeq 0.29$), corresponds to the
yielding point. Until this point the honeycomb like structure
is preserved, but beyond it the plastic deformation sets in.
We note that for $\epsilon_{c_{1}}<\epsilon<\epsilon_{c_{2}}$
the system is actually in a metastable state, where the
plastic deformation is delayed. Under the long wavelength
perturbations, vacancy defects and high ambient temperature
$\epsilon_{c_{2}}$ approaches to $\epsilon_{c_{1}}$. In fact,
our further molecular dynamics simulations show that
$\epsilon_{c_{2}} \rightarrow$ 0.17 at 300K and to 0.16 at
600K. In the presence of periodically repeating F-vacancy and
C+F-divacancy, the value of $\epsilon_{c_{2}}$ is also
lowered to 0.21 and 0.14, respectively. Apart from phonon
instability occurring at high $\epsilon$, the band gap is
strongly affected under uniform expansion. In
Fig.~\ref{figs5}(b) we show the variation of LDA and $GW_0$-
corrected band gaps under uniform expansion. The LDA gap
slightly increases until $\epsilon=0.05$ and then decreases
steadily with increasing $\epsilon$. The $GW_0$-corrected
band gap essentially decreases with increasing strain. For
example, its value decreases by 38\% for
$\epsilon=0.20$.
\section{Conclusions}
Present analysis of fluorinated graphenes shows that
different C$_n$F structures can form at different level of
F coverage. Calculated properties of these structures, such
as lattice parameter, $d_{CC}$ distance, band gap, density
of states, work function, in plane stiffness $C$, Poisson's
ratio and surface charge, are shown to depend on $n$ or
coverage $\Theta$. Relevant data reported in various
experiments do not appear to agree with the properties
calculated any one of the stable C$_n$F structures. This
finding lets us to conclude that domains of various C$_n$F
structures can form in the course of the fluorination of
graphene. Therefore, the experimental data may reflect a
weighted average of diverse C$_n$F structures together with
extended defects in grain boundaries. In this respect,
imaging of fluorinated graphene surfaces by scanning
tunneling and atomic force microscopy, as well as x-ray
photoemission spectroscopy is expected to shed light on the
puzzling inconsistency between theory and experiment.
Finally, our results show a wide range of interesting
features of C$_n$F structures. For example, a perfect CF
structure as described in Fig.~\ref{fig2} is a stiff,
nonmagnetic wide band gap nanomaterial having substantial
surface charge, but attains significant local magnetic moment
through F-vacancy defects. Moreover, unlike graphane, half
fluorinated graphene with only one side
fluorinated is found to be stable, which can be further
functionalized by the adsorption of adatoms to other side.
For example, hydrogen atoms adsorbed to other side attain
positive charge and hence permanent transversal electric
field, which can be utilized to engineer electronic
properties.
\section{ACKNOWLEDGMENTS}
This work is supported by TUBITAK through Grant No:108T234.
Part of the computational resources has been provided by
UYBHM at ITU through grant Grant No. 2-024-2007. We thank the
DEISA Consortium (www.deisa.eu), funded through the EU FP7
project RI-222919, for support within the DEISA Extreme
Computing Initiative. S. C. acknowledges the partial support
of TUBA, Academy of Science of Turkey. The authors would also
like to acknowledge the valuable suggestions made by D. Alfe.
|
\section{Introduction and summary of results}
Black hole physics suggests that gravitational theories are holographic
and indeed over the last 15 years we have witnessed remarkable progress in
uncovering holographic relations, mainly between $(d+2)$-dimensional gravitational theories
with negative cosmological constant and $(d+1)$-dimensional local quantum
field theories. The case of asymptotically flat spacetimes remains elusive
and indeed the structure of the asymptotic solutions
\cite{BS,BS2,deHaro:2000wj} and the divergences of the
on-shell action suggest \cite{Skenderis:2002wp} that
if there is a dual theory it is nonlocal, see also \cite{Papadimitriou:2010as}.
In a recent paper \cite{Bredberg} (see
\cite{Andy, Fouxon:2008tb,Bhattacharyya:2008kq,Eling:2009pb, Padmanabhan, Damour1, Damour2, Thorne} for earlier
relevant work),
a remarkable relation between
incompressible non-relativistic fluids in $(d+1)$ dimensions
satisfying the Navier-Stokes equation and
$(d+2)$-dimensional Ricci
flat metrics was found. The construction of \cite{Bredberg} starts
from considering the portion of Minkowski spacetime between a flat
hypersurface $\Sigma_c$, given by the equation $X^2 - T^2 = 4 r_c$, and
its future horizon ${\cal H}^+$, the null surface $X=T$. Then the effects of finite perturbations
of the extrinsic curvature
of $\Sigma_c$, with the intrinsic metric $\g_{ab}$ kept fixed, were studied and it was found that
a regular Ricci flat metric exists provided that the Brown-York
stress tensor on $\Sigma_c$ is that of an incompressible Navier-Stokes
fluid. More precisely, they work in a hydrodynamic non-relativistic
limit and construct the bulk metric up to third order in the
hydrodynamic expansion. Apart from its intrinsic value, this
relationship provides a potential example of a holographic duality
involving a flat spacetime and as such it should be further explored.
The aim of this paper is two-fold. Firstly, we would like to provide
a systematic construction of the bulk solution to all orders in the hydrodynamic
expansion and secondly we would like to extract the physical properties of the
dual fluid in order to obtain clues about the nature of the dual theory.
Let us summarise our results. First recall that in the hydrodynamic limit
essentially any $(d+1)$-dimensional fluid is governed by
the incompressible Navier-Stokes equation
\[ \label{NS}
E_{i} \equiv \D_\t v_i + v^j\D_jv_i-\eta\D^2 v_i +\D_i P = 0, \qquad
E \equiv \D_i v^i = 0, \quad (i = 1, \ldots, d),
\]
where $v^i$ is the velocity field, $P$ is a pressure fluctuation about the
background value and $\eta$ is the kinematic viscosity.
This equation has a scaling symmetry and admits a one-parameter family of
solutions
\[
v_i^{(\ep)}(\t,\vec{x}) = \ep v_i(\ep^2\t,\ep\vec{x}), \qquad P^{(\ep)}(\t,\vec{x})=\ep^2 P(\ep^2\t,\ep \vec{x}).
\]
If one considers now higher-derivative corrections to $E_{i}$ and
$E$, these are naturally organised according to their scaling with
$\ep$.
Our first main result is an algorithmic construction of a
regular Ricci-flat metric corresponding to a fluid satisfying
(\ref{NS}) with specific higher-derivative corrections. We begin with the
observation that
the metric up to order $\ep^2$ with constant velocity and pressure fields
is actually flat space in disguise: it can be
obtained from the Rindler metric by a linear coordinate transformation
combining a boost which introduces a constant velocity field
$v^i$ with a simple shift and rescaling of the radial and time coordinates that
introduces a constant pressure function $P$. In fact, these transformations
are the most general infinitesimal diffeomorphisms that preserve the
induced metric at $r = r_c$ and lead to a Brown-York stress tensor which
is of the form of a fluid in equilibrium. (Recall that the Brown-York stress
tensor for Rindler spacetime is of that form.)
To extend the solution to next order in the hydrodynamic expansion,
we follow \cite{Bhattacharyya:2008jc} and promote
the velocity field and the pressure function to be
spacetime dependent quantities, subject to the requirement that the
Einstein equations hold up to order $\ep^3$. To satisfy this
requirement one needs to introduce new terms of order $\ep^3$ in the
metric and we find that (\ref{NS})
holds. This reproduces the metric
given in \cite{Bredberg} (up to a choice of gauge for the dual fluid).
Next we prove inductively that
Einstein equations can in fact be satisfied to arbitrarily high order in $\ep$,
by adding appropriate terms to the metric and modifying the
Navier-Stokes equation and the incompressibility condition by specific higher
derivative corrections. In particular, the incompressibility
condition $E$ is modified at even powers of $\ep$ while the
Navier-Stokes equation $E_{i}$ is modified at odd powers of $\ep$.
We carried out this procedure explicitly to order $\ep^5$ and obtained
the resultant metric through to order $\ep^5$, given in
(\ref{g3})-(\ref{g41})-(\ref{g42})-(\ref{g5})
as well as the leading order corrections to $E$ and $E_{i}$, given in
(\ref{final_I}) and (\ref{final_NS}), respectively.
The dual fluid has a number of interesting and unconventional features.
Firstly, in equilibrium it has zero energy density and positive nonzero
pressure. In the hydrodynamic regime, the stress tensor receives
dissipative corrections,
the explicit form of which is given to order $\ep^5$ in
(\ref{T3}), (\ref{T4}) and (\ref{T5}). As a consequence of the (bulk)
Hamiltonian constraint, the fluid stress tensor satisfies
the quadratic constraint,
\[ \label{qua-con}
d T_{ab}T^{ab} = T^2.
\]
This is an interesting constraint on the dual theory which will be discussed
further below.
Since the bulk solution is constructed using a non-relativistic hydrodynamic
expansion in $\ep$, the fluid stress tensor appears to be
non-relativistic. In turns out however that the complete answer, including
dissipative terms, can be obtained from a relativistic hydrodynamic stress
tensor upon using the non-relativistic expansion in $\ep$. The rather
complicated expressions obtained by direct computation actually originate from
a much simpler relativistic stress tensor. The information
encoded in the stress tensor can then be expressed in terms of
a few transport coefficients. These considerations indicate also that
the underlying holographic dual theory should be relativistic.
Recall that a relativistic fluid stress tensor takes the general form
\[
T_{ab} = \rho u_a u_b + p h_{ab} + \Pi^\perp_{ab}, \qquad u^a\Pi^\perp_{ab}=0,
\]
where $\rho$ and $p$ are the energy density and pressure of the fluid in
the local rest frame, $u^a$ is the relativistic fluid velocity (normalised such that $u_au^a=-1$),
and $h_{ab} = \g_{ab}+u_au_b$ is the induced metric on surfaces orthogonal
to the fluid velocity. The term $\Pi^\perp_{ab}$ encodes dissipative
corrections and may be expanded in gradients of the fluid velocity.
When the equilibrium energy density is nonzero, one may show that
the Landau-Lifshitz prescription \cite{Landau} implies that the
energy density $\rho$ does not receive
any gradient corrections. This is in stark contrast to the case in hand
where the equilibrium fluid has zero energy density and the energy density becomes
nonzero and negative due to gradient terms. In fact the quadratic
constraint (\ref{qua-con}) uniquely determines $\rho$ in terms of the dissipative terms.
The leading order result for $\rho$ is given in (\ref{rho_lead}).
From the dissipative terms, $\Pi^\perp_{ab}$, one can read off the transport
coefficients. The relativistic corrections are organised according to
the number of derivatives they contain. Each of these terms can then be
expanded in the non-relativistic hydrodynamic expansion.
The $\ep$ expansion may increase the number of derivatives but it never
decreases them. This means that one can unambiguously extract all
transport coefficients but one will have to work to sufficiently high order
in $\ep$ in order to obtain all transport coefficients. At first order
in derivatives, there is only one possible term\footnote{For relativistic fluids in general
there would also be a bulk viscosity term.
For fluids with zero equilibrium energy density, this term vanishes
as a consequence of the equations of motion (see section \ref{Section:Hydro}).}
and from this the kinematic shear viscosity is obtained.
If we identify the entropy density of the fluid using the entropy of
the Rindler horizon, the ratio $\eta/s$ takes the universal value,
\begin{equation}
\eta/s = 1/(4 \pi).
\end{equation}
At second order, there are six transport coefficients (for a fluid in a flat background)
and our computation to order $\epsilon^5$ determines four of them, see
(\ref{PiT}) and (\ref{transport_coefficients}). The products of these coefficients with
the (background) pressure are pure numbers and one gets the following relations
\begin{equation}
2 c_1 p = c_2 p= c_3 p = c_4 p = -4.
\end{equation}
These results encode all the information contained in the dissipative parts of the
stress tensor, up to this order.
Let us now move to the non-dissipative part,
\begin{equation}\label{nondis}
T_{ab}^{\rm non{-}dissipative} = p h_{ab}.
\end{equation}
Note that this satisfies by itself the quadratic constraint.
One may ask whether there is a Lagrangian model that leads
to this stress tensor. It turns out that such a model indeed
exists and is given by\footnote{Interestingly, this
action has also been investigated
in the context of dark energy, see \cite{Niayesh1, Niayesh2}.}
\[
S = \int \d^{d+1}x \sqrt{-\g} \sqrt{-(\D\phi)^2}.
\]
This describes an ideal fluid
with $u_a = \D_a \phi/\sqrt{-(\D\phi)^2}$ and stress tensor
given by (\ref{nondis}) with $p=\sqrt{-(\D\phi)^2}$. The equilibrium configuration
corresponds to the solution $\phi=\t$ and near-equilibrium configurations are obtained by looking at small
fluctuations around this solution. The background solution breaks the relativistic invariance and the expansion around this solution
leads to a non-relativistic nonlocal action (since it contains an infinite number of derivatives).
Thus, this simple scalar theory (and its generalisations, see \eqref{gen-mod}) reproduces the non-dissipative
part of the stress tensor and incorporates many of the features ones expects
from the holographic dual theory.
\bigskip
The plan of the paper is as follows. In section \ref{eq-configurations}, we discuss a general class of flat metrics admitting a Rindler horizon
for which the Brown-York stress tensor on the constant-radius slice $\Sigma_c$ outside the horizon is that of a perfect fluid with zero energy density. These equilibrium
backgrounds are used in section \ref{Se:seed} as seed metrics to construct corresponding near-equilibrium configurations. In section \ref{Section_algorithm}, we
give the algorithm for systematically constructing the bulk solution to arbitrary order in the hydrodynamic expansion, while in section \ref{gauge-choice}, we explain
the gauge-fixing conditions imposed on the fluid. In section \ref{explicit}, we give explicit results for the metric and stress tensor up to order $\ep^5$, and
in section \ref{Section:Hydro}, these results are interpreted in terms of transport coefficients for the dual fluid. Section \ref{Section:Models} proposes a simple
dual Lagrangian which captures key properties of the fluid, and in section \ref{Discussion}, we discuss our results. Finally, in the three appendices we present some of the technical results used in the main text.
In appendix \ref{app:radialgauge}, we discuss the choice of radial gauge;
in appendix \ref{app:no_rho}, we
derive the class of infinitesimal diffeomorphisms preserving the equilibrium form of the Brown-York stress tensor; and in appendix \ref{hydro_app}, we derive the general form
of the hydrodynamic expansion to second order in gradients for a relativistic fluid with vanishing equilibrium energy density.
\section{Equilibrium configurations}
\label{eq-configurations}
Let us first construct a class of Riemann-flat metrics which have the following properties: (i) they admit a co-dimension one hypersurface $\Sigma_c$ of flat induced metric
\[
\g_{ab}\d x^a\d x^b=-r_c\d\t^2+\d x_i\d x^i,\label{indmetric}
\]
where the speed of light $\sqrt{r_c}$ is arbitrary,
$x^\mu = (r, x^a)$ and $x^a=(\t,x^i)$ with $i=1,{\ldots},d$
(unless otherwise noted, Latin indices are raised with $\gamma^{ab}$ and Greek indices with $g^{\mu\nu}$);
(ii) the Brown-York stress tensor \cite{Brown:1992br} on $\Sigma_c$, given by\footnote{From here onwards we will set $16 \pi G = 1$.
We also assume the validity of classical gravity, i.e., higher-order curvature and quantum corrections are suppressed.}
\[
T_{ab}=\frac{1}{8\pi G}(K \gamma_{ab}-K_{ab}),
\]
where $K_{ab}$ is the extrinsic curvature of $\Sigma_c$, takes the form of a perfect fluid stress tensor,
\[
T_{ab} = \rho u_a u_b+ p h_{ab}\label{stresstensor},
\]
and finally; (iii) they are stationary with respect to $\partial_\tau$ and homogeneous in the $x^i$ directions.
One such metric with these properties is
\[
\label{Rindler}
\d s^2 = \bar g_{\mu\nu}dx^\mu dx^\nu = -r\d\t^2+2\d\t\d r + \d x_i\d x^i,
\]
which describes flat space in ingoing Rindler coordinates. (Upon setting $\t=2\ln(X+T)$ and $4r=X^2-T^2$, we recover $\d s^2=-\d T^2+\d X^2$.)
The induced metric $\g_{ab}$ on the surface $\Sigma_c$ defined by $r=r_c$ then has the form \eqref{indmetric}, while the Brown-York stress tensor on $\Sigma_c$ has the form \eqref{stresstensor}, where $u_\tau = -r_c^{1/2}$, $u_i=0$, $\rho = 0$ and $p = r_c^{-1/2}$. The Rindler horizon is at $r=0$.
We may now obtain further metrics satisfying conditions (i), (ii) and (iii) by considering the action of diffeomorphisms on \eqref{Rindler}.
Since we wish to obtain a connected thermodynamical state space, it is sufficient to consider only diffeomorphisms that are continuously connected to the identity, i.e., those obtained by exponentiating infinitesimal diffeomorphisms.
In appendix \ref{app:no_rho}, we show that there are only two infinitesimal diffeomorphisms yielding metrics satisfying conditions (i), (ii) and (iii), after fixing the radial gauge as described in appendix \ref{app:radialgauge}.
Exponentiating these, we obtain the following two finite diffeomorphisms.
The first is a constant boost $\beta_i$, given by
\[
\sqrt{r_c}\t \rightarrow \g\sqrt{r_c}\t-\g\beta_i x^i, \qquad
x^i \rightarrow x^i -\g\beta^i\sqrt{r_c}\t+(\g-1)\frac{\beta^i\beta_j}{\beta^2}x^j,\label{t1}
\]
where $\g=(1-\beta^2)^{-1/2}$ and $\beta_i \equiv r_c^{-1/2} v_i$. The second is a constant linear shift of $r$ and associated re-scaling of $\tau$,
\[
r \rightarrow r-r_h, \qquad \t \rightarrow (1-r_h/r_c)^{-1/2}\t.\label{t2}
\]
Note that this latter transformation shifts the position of the horizon to $r=r_h$; as we are only interested in the case where $\Sigma_c$ is outside the horizon we will restrict $r_h < r_c$.
Since these transformations are both linear in the coordinates, one can apply them in either order on Rindler spacetime. One obtains
\begin{align}
\d s^2 &= \frac{\d\tau^2}{1-v^2/r_c}\(v^2 - \frac{r-r_h}{1-r_h/r_c}\)+\frac{2\gamma}{\sqrt{1-r_h/r_c}}\d\tau \d r - \frac{2\gamma v_i}{r_c \sqrt{1-r_h/r_c}}\d x^i \d r \nn \\[1ex]
&\quad + \frac{2v_i}{1-v^2/r_c}\( \frac{r-r_c}{r_c-r_h}\)\d x^i \d\tau + \(\delta_{ij}- \frac{v_iv_j}{r_c^2 (1-v^2/r_c)} \(\frac{r-r_c}{1-r_h/r_c}\)\)\d x^i \d x^j\, .\label{eq:metric}
\end{align}
While this metric appears complicated, it is in reality simply flat space written in a complicated coordinate system.
Nevertheless, the Brown-York stress tensor on $\Sigma_c$ has the desired form \eqref{stresstensor}, where the energy density, pressure and relativistic fluid velocity are given by
\[
\rho = 0,\qquad p = \frac{1}{\sqrt{r_c-r_h}} ,\qquad u^a = \frac{1}{\sqrt{r_c-v^2}}(1,v_i) \label{pressequil}\, .
\]
In particular, the energy density $\rho$ vanishes.
We stress that there is no flat solution continuously connected to Rindler spacetime which generates a nonzero equilibrium energy density while at the same time preserving the form of the induced metric on $\Sigma_c$, as shown in appendix \ref{app:no_rho}.
In connection with this, note that the inherent ambiguity in the definition of the Brown-York stress tensor \cite{Brown:1992br},
which gives the freedom to send $T_{ab}\rightarrow T_{ab}+C\g_{ab}$ for some constant $C$, produces only a trivial constant shift in the background energy density and pressure given by $\rho \rightarrow \rho-C$ and $p\rightarrow p+C$. This ambiguity has no dynamical consequences, however, and so we will suppress it in the remainder of this paper by setting $C=0$. Furthermore, since the induced metric $\g_{ab}$ is flat, there are no additional ambiguities involving the intrinsic curvature of $\Sigma_c$.
Finally, let us consider the thermodynamic properties of the metric \eqref{eq:metric}.
First, the Rindler horizon is a Killing horizon and has a constant Unruh temperature given by
\[
T = \frac{\kappa}{2\pi} = \frac{1}{4\pi \sqrt{r_c-r_h}},
\]
where the surface gravity $\kappa$ is defined as $\xi^\nu \nabla_\nu \xi^\mu = \kappa \xi^\mu$, where $\xi = \frac{1}{\sqrt{r_c -v^2}} (\partial_\tau + v_i \partial_i)$ is normalised so that $\xi_a \xi^a = -1$ on $\Sigma_c$. If we then introduce an entropy density associated with the Rindler horizon identical to the black hole entropy density $s=1/4G$, we note the thermodynamic relation\footnote{The relation \eqref{eq:stp} may also be derived by evaluating Komar integrals both at the horizon $\Sigma_h$ and at $\Sigma_c$. Since the solution is stationary, homogeneous in $x^i$ and Ricci flat, the Komar integrals do not depend on the radius of evaluation and one can relate $sT$ defined at the horizon to $p$ defined on $\Sigma_c$ using standard manipulations, see e.g., \cite{Bardeen:1973gs,Townsend:1997ku}.}
\[
sT = p\, .\label{eq:stp}
\]
This further implies that two neighbouring equilibrium configurations are related by the Gibbs-Duhem relation
\[
s \delta T = \delta p \,.
\]
As a final point of interest, we note that the region $r_h \leq r \leq r_h + (1-r_h/r_c)v^2$ is an ergoregion with respect to the $\tau$-translations, which may lead to effects analogous to superradiance.
\section{Seed metric for near-equilibrium configurations}
\label{Se:seed}
Let us now change gears and discuss how to construct near-equilibrium configurations. We would like to obtain a solution of the vacuum Einstein equations parameterised by a velocity field $v_i$ and a pressure field $p$ which are local slowly varying functions of the coordinates $x^a$,
while preserving the form of the induced metric on $\Sigma_c$.
Working perturbatively in $\ep$, in the present section we show how to construct this solution up to order $\ep^2$.
In the following section, we then show how to extend this `seed' solution through to all orders in $\ep$.
Clearly, if we promote $v_i$ and $p$ to depend arbitrarily on the coordinates $x^a$, the metric \eqref{eq:metric} is no longer an exact solution of the vacuum Einstein equations. However, treating $v_i = v_i^{(\ep)}(\t,\vec{x}) $ and $p =r_c^{-1/2} + r_c^{-3/2}P^{(\ep)}(\t,\vec{x})$ as small fluctuations around the background $v_i = 0$, $p=r_c^{-1/2}$ in the hydrodynamic limit
\[
v_i^{(\ep)}(\t,\vec{x}) = \ep v_i(\ep^2\t,\ep\vec{x}), \qquad P^{(\ep)}(\t,\vec{x})=\ep^2 P(\ep^2\t,\ep \vec{x}),
\]
upon expanding \eqref{eq:metric} (noting that $r_h=2P^{(\ep)}+O(\ep^4)$) we obtain
\begin{align}
\label{seed}
\d s^2 &= -r \d\t^2+2\d\t\d r+\d x_i\d x^i \nn\\
&\quad -2\(1-\frac{r}{r_c}\)v_i\d x^i \d \t-\frac{2v_i}{r_c}\d x^i \d r \nn\\
&\quad +\(1-\frac{r}{r_c}\)\Big[(v^2+2P)\d\t^2+\frac{v_iv_j}{r_c}\d x^i\d x^j\Big]+\(\frac{v^2}{r_c}+\frac{2P}{r_c}\)\d\t\d r
+O(\ep^3),
\end{align}
where each successive line is of one order higher in the $\epsilon$ expansion. (Here, and in the following, we drop the superscripts on $P^{(\ep)}$ and $v^{(\ep)}_i$ for simplicity.) This metric by construction
preserves the induced metric $\g_{ab}$ on $\Sigma_c$. Moreover, it solves Einstein's equations to $O(\ep^3)$, provided one additionally imposes incompressibility, i.e., $\D_iv^i = O(\ep^3)$. The metric \eqref{seed} reproduces the solution found in \cite{Bredberg} through to $\ep^2$ order.
As shown in \cite{Bredberg}, the Brown-York stress tensor on $\Sigma_c$ for the seed solution \eqref{seed} is
\begin{align}
\label{seed_stress}
T_{ab}\d x^a \d x^b &= \frac{\d \vec{x}^2}{\sqrt{r_c}}
- \frac{2 v_i}{\sqrt{r_c}}\,\d x^i\d \t
+\frac{v^2}{\sqrt{r_c}}\,\d\t^2 + r_c^{-3/2}\Big[P\delta_{ij}+v_iv_j-2r_c \D_i v_j\Big]\d x^i\d x^j +O(\epsilon^3)\, .
\end{align}
It is interesting to compare this result with the hydrodynamic stress tensor for a relativistic fluid with vanishing equilibrium energy density.
Up to first order in fluid gradients, this takes the form\footnote{As noted previously, and as we will discuss in section \ref{Section:Hydro}, for a fluid with vanishing equilibrium energy density the relativistic divergence $\K=\D_au^a$ vanishes at first order in gradients and so the usual bulk viscosity term is absent.}
\[
\label{std_hydroT}
T_{ab} = \rho u_a u_b + p h_{ab} -2 \eta \K_{ab} + O(\D^2) ,
\]
where $\K_{ab} = h_a^c h_b^d\D_{(c}u_{d)}$ is the extrinsic curvature of surfaces orthogonal to the fluid velocity\footnote{We reserve the symbol $K_{ab}$ for the extrinsic curvature of $\Sigma_c$ itself.} and plays the role of the fluid shear.
Expanding this hydrodynamic stress tensor in $\ep$ to order $\ep^2$, one finds that one can indeed reproduce \eqref{seed_stress} upon setting
\[
\rho = 0 + O(\epsilon^3),\qquad p = \frac{1}{\sqrt{r_c}} + \frac{P}{r_c^{3/2}}+ O(\epsilon^3)\, ,\qquad \eta = 1\, .
\]
In the following section, we will now describe the algorithm to extend our seed solution \eqref{seed} to arbitrarily high order in $\epsilon$,
with the pressure and velocity perturbations obeying the incompressible Navier-Stokes equations with higher order corrections. Note that, if desired,
spacetimes describing fluctuations around a background with nonzero velocity and with a Rindler horizon located at
arbitrary $r=r_h$ may then be found by applying the finite diffeomorphisms \eqref{t1} and \eqref{t2} to the resulting solution.
\section{Constructing the solution to all orders}
\label{Section_algorithm}
Let us begin by assuming we know the bulk metric through order $\ep^{n-1}$.
The first nonvanishing components of the Ricci tensor appear therefore at order $\ep^n$.
We now wish to add a new term $g^\n_{\mu\nu}$ to the metric at order $\ep^n$ so as to extend our solution to order $O(\ep^{n+1})$. In addition to adding a new term to the metric, we must also consider the effect of gauge transformations $\xi^{\n\mu}$ and field redefinitions $\delta v_i^\n$ and $\delta P^\n$ also at order $\ep^n$.
Noting that
\[
\D_r \sim \ep^0, \qquad \D_i \sim \ep^1, \qquad \D_\t \sim \ep^2,
\]
adding a new term $g^\n_{\mu\nu}$ at order $\ep^n$ to the bulk metric produces a change in bulk Ricci tensor
\noindent at the same order given by
\begin{align}
\label{linearRicci}
\delta R_{rr}^\n &= -\frac{1}{2}\D_r^2 g^\n_{ii}, \nn\\
\delta R_{ij}^\n &= -\frac{1}{2}\D_r(r\D_r g^\n_{ij}), \nn\\
\delta R_{\t i}^\n &= -r\delta R_{ri}^\n = -\frac{r}{2}\D_r^2 g_{\t i}^\n, \nn\\
\delta R_{\t \t}^\n &=-r \delta R_{r \t}^\n = -\frac{r}{4} \(\D_r(r g^\n_{rr})+2\D_r g_{r\t}^\n-\D_r g_{ii}^\n+2\D_r^2 g_{\t\t}^\n\),
\end{align}
where we write $g_{ii}^\n \equiv \delta^{ij} g_{ij}^\n$ and $\delta R_{ii}^\n \equiv \delta^{ij}\delta R_{ij}^\n$.
Our goal then is to find a $g_{\mu\nu}^\n$ which cancels out the part of the Ricci tensor at order $\ep^n$ arising from the pre-existing metric up to order $\ep^{n-1}$.
(In fact, as we will see shortly, the boundary conditions are such that this requirement leads, after gauge-fixing, to a unique solution for
$g_{\mu\nu}^\n$.)
The full Ricci tensor at order $\ep^n$ must therefore satisfy
\[
\label{EinsteinEq}
R^{(n)}_{\mu\nu} = \delta R_{\mu\nu}^\n + \hat{R}_{\mu\nu}^\n = 0,
\]
where $\hat{R}_{\mu\nu}^\n$ denotes the part of the Ricci tensor at order $\ep^n$ arising from the metric up to order $\ep^{n-1}$.
Inspecting the equations for $\delta R_{\mu\nu}^\n$ above, we see that such a cancellation will only be possible provided the following integrability conditions are satisfied:
\[
\label{int_condns}
0 = \D_r(\hat{R}_{ii}^\n-r\hat{R}_{rr}^\n) -\hat{R}_{rr}^\n, \qquad
0 = \hat{R}_{\t a}^\n+r\hat{R}_{r a}^\n.
\]
To verify that these conditions are indeed satisfied, we
evaluate the Bianchi identity at order $\ep^n$ yielding\footnote{Note that the Ricci tensor $R_{\mu\nu}$ is itself of order $\ep^n$ (as we have already solved the vacuum Einstein equations at order $\ep^{n-1}$), so all the remaining covariant derivatives, metrics and inverse metrics appearing in the Bianchi identity are those of the background Rindler metric.}
\begin{align}
\label{Bianchi_id}
0 &= \D_r(\hat{R}_{ii}^\n-r\hat{R}_{rr}^\n) -\hat{R}_{rr}^\n, \nn\\
0 &= \D_r(\hat{R}_{\t a}^\n+r\hat{R}_{r a}^\n) \quad \Rightarrow \quad \hat{R}_{\t a}^\n+r\hat{R}_{r a}^\n = f_a^\n(\t,\vec{x}).
\end{align}
The integrability conditions are therefore satisfied provided the arbitrary function $f_a^\n(\t,\vec{x})$ vanishes.
Evaluating the Gauss-Codazzi identity on $\Sigma_c$ at order $\ep^n$, we find
\[
\label{Gauss_Codacci}
\nabla^bT_{ab}\big|^\n_{\Sigma_c} = [2\nabla^b(K\g_{ab}-K_{ab})]^\n = [-2R_{a\mu}N^\mu]^\n = -\frac{2}{\sqrt{r_c}}(\hat{R}_{\t a}^\n+r_c\hat{R}_{r a}^\n)=-\frac{2}{\sqrt{r_c}}f_a^\n(\t,\vec{x}).
\]
Thus, imposing the conservation of the Brown-York stress tensor on $\Sigma_c$ at order $\ep^n$ enforces the vanishing of $f_a^\n(\t,\vec{x})$, ensuring the validity of the integrability conditions \eqref{int_condns}.
From the perspective of the dual fluid, at order $\ep^3$ (i.e., the first order beyond the seed solution), this constraint reduces to the exact Navier-Stokes equation. At subsequent even orders in $\ep$, this step amounts to adding corrections to the incompressibility condition, while at higher odd orders it amounts to adding corrections to the Navier-Stokes equation.
It is interesting to note
that evaluating the Hamiltonian constraint on $\Sigma_c$ yields an exact and nontrivial constraint on the Brown-York stress tensor of the dual fluid, namely
\[
\label{Ham_constraint}
d T_{ab}T^{ab} = T^2.
\]
Applying this constraint alone to equilibrium fluid configurations, one finds that either the equilibrium energy density vanishes $\rho_{\mathrm{eqm}}=0$ (as is the case here) or else is given by $\rho_{\mathrm{eqm}}=\left(-2d/(d-1)\right) p$, where $p$ is the equilibrium pressure. It would be interesting to see if this latter solution branch could be obtained through a modification of the present scenario. We will discuss further the implications of the constraint \eqref{Ham_constraint} for the dual hydrodynamics in section \ref{Section:Hydro}.
We may now proceed to use \eqref{linearRicci} to identify a particular solution $\tilde{g}^\n_{\mu\nu}$ that satisfies the Einstein equation \eqref{EinsteinEq} at order $\ep^n$. One such particular solution is
\begin{align}
\tilde{g}_{r\mu}^\n &= 0,\nn\\
\tilde{g}_{\t\t}^\n &= \beta_1^\n(\t,\vec{x})+(1-r/r_c)\beta_2^\n(\t,\vec{x})+\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' (\hat{R}_{ii}^\n-r\hat{R}_{rr}^\n-2\hat{R}_{r\t}^\n),\nn\\
\tilde{g}_{\t i}^\n &= \beta_{3i}^\n(\t,\vec{x})+(1-r/r_c)\beta_{4i}^\n(\t,\vec{x})-2\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' \hat{R}_{ri}^\n,\nn\\
\tilde{g}_{ij}^\n&=\beta_{5ij}^\n(\t,\vec{x})+\ln(r/r_c)\beta_{6ij}^\n(\t,\vec{x})-2\int^{r_c}_r\d r' \frac{1}{r'}\int^{r'}_{r_*}\d r'' \hat{R}_{ij}^\n.
\end{align}
Here, we have chosen limits so that the integrals vanish for $r=r_c$, facilitating the evaluation of boundary conditions to follow shortly.
We will leave the lower limit $r_*$ appearing in the inner integral for $\tilde{g}_{ij}^\n$ arbitrary for now.
Note however that the $rr$ Einstein equation, in combination with the Bianchi identity \eqref{Bianchi_id}, fixes the trace $\beta_{6ii}^\n=\delta^{ij}\beta_{6ij}^\n$ to be
\[
\label{beta_constraint}
\beta_{6ii}^\n(\t,\vec{x}) = 2r_*(\hat{R}_{ii}^\n-r_*\hat{R}_{rr}^\n)\big|_{r=r_*}.
\]
Allowing now for gauge transformations $\xi^{\n\mu}(r,\t,\vec{x})$ at order $\ep^n$, as well as field redefinitions $\delta v_i^\n(\t,\vec{x})$ and $\delta P^\n(\t,\vec{x})$ also at this order, the above solution generalises to
\begin{align}
\label{gsoln}
g_{rr}^\n &= 2\D_r\xi^{\n\t}, \nn\\
g_{r\t}^\n &= -r\D_r\xi^{\n\t}+\D_r\xi^{\n r}+\frac{1}{r_c}\delta P^\n,\nn\\
g_{ri}^\n &= \D_r\xi_i^\n-\frac{1}{r_c}\delta v_i^\n,\nn\\
g_{\t\t}^\n &= \tilde{g}_{\t\t}^\n-\xi^{\n r}+(1-r/r_c)2\delta P^\n,\nn\\
g_{\t i}^\n &= \tilde{g}_{\t i}^\n-2(1-r/r_c)\delta v_i^\n, \nn\\
g_{ij}^\n &= \tilde{g}_{ij}^\n.
\end{align}
Here, to obtain the terms involving $\delta v_i^\n$ and $\delta P^\n$ it is sufficient to look at
how $v_i$ and $P$ appear linearly in the seed solution \eqref{seed}, since $\delta v_i^\n \sim \delta P^\n \sim \ep^n$.
Note that the $g_{\mu\nu}^\n$ above is still a solution of Einstein's equations, since the additional terms
coming from field redefinitions and gauge transformations do not contribute to $\delta R_{\mu\nu}^\n$, as we see from \eqref{linearRicci}.
To fix the gauge, we will impose $g^\n_{r\mu}=0$, for all $n>2$. This choice is convenient since we do not have boundary conditions at $\Sigma_c$ for these metric components, unlike for the $g^\n_{ab}$.
The $g_{r\mu}$ components of the full metric are then simply those of the seed solution, i.e., to all orders in $\ep$,
\[
\label{gr_components}
g_{rr} = 0, \qquad g_{r\t} =1+\frac{v^2}{2r_c}+\frac{P}{r_c} ,\qquad g_{ri} = -\frac{v_i}{r_c}.
\]
Note this gauge choice is fully consistent with our considerations in appendix \ref{app:radialgauge} (cf.~equation \eqref{gauge1}).
To impose this gauge choice then, we must fix
\begin{align}
& \xi^{\n r} = (1-r/r_c)\delta P^\n +\tilde{\xi}^{\n r}(\t,\vec{x}), \qquad \xi^{\n\t} = \tilde{\xi}^{\n\t}(\t,\vec{x}),\nn\\[1ex]
&\qquad\qquad \xi^\n_i = -(1-r/r_c)\delta v_i^\n+\tilde{\xi}^\n_i(\t,\vec{x}),
\end{align}
where the $\tilde{\xi}^{\n\mu}(\t,\vec{x})$ parameterise the residual gauge freedom.
Substituting this back into \eqref{gsoln}, we find
\begin{align}
\label{newgsoln}
g_{r\mu}^\n &= 0, \nn\\
g_{\t\t}^\n &= \beta_1^\n-\tilde{\xi}^{\n r}+(1-r/r_c)(\beta_2^\n+\delta P^\n)+\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' (\hat{R}_{ii}^\n-r\hat{R}_{rr}^\n-2\hat{R}_{r\t}^\n),\nn\\
g_{\t i}^\n &= \beta_{3i}^\n+(1-r/r_c)(\beta_{4i}^\n-2\delta v_i^\n)-2\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' \hat{R}_{ri}^\n, \nn\\
g_{ij}^\n &= \beta_{5ij}^\n+\ln(r/r_c)\beta_{6ij}^\n-2\int^{r_c}_r\d r' \frac{1}{r'}\int^{r'}_{r_*}\d r'' \hat{R}_{ij}^\n.
\end{align}
Imposing the boundary condition $g^\n_{ab}=0$ on $\Sigma_c$ (so that the induced metric $\g_{ab}$ remains fixed) then fixes
\[
\label{beta_fix}
\beta_1^\n(\t,\vec{x}) = \tilde{\xi}^{\n r}(\t,\vec{x}), \qquad \beta_{3i}^\n(\t,\vec{x})=0, \qquad \beta_{5ij}^\n(\t,\vec{x})=0.
\]
To fix the trace-free part of $\beta_{6ij}^\n$, we require that the metric must be regular at the future horizon $r=0$, or equivalently analytic, at each order $\ep^n$.
To achieve this, it is useful to choose the lower limit $r_*$ in the inner integral for $g_{ij}^\n$ in \eqref{newgsoln} to be zero.
Since $\hat{R}_{ij}^\n$ will turn out to be a polynomial in $r$, by setting $r_*=0$ we ensure that the outer integral does not generate any additional logarithmic terms beyond those associated with the $\beta_{6ij}^\n$. (Moreover, the trace part of $\beta_{6ij}^\n$ is set to zero through \eqref{beta_constraint}.)
Regularity on the future horizon at order $\ep^n$ then requires\footnote{A possible caveat here is that we impose
order by order regularity but the logarithmic terms could yield a finite value on the future horizon when resummed to all orders in $\ep$. We thank Bob Wald for discussion on this point.}
that these remaining logarithmic terms vanish:
\[
\beta_{6ij}^\n(\t,\vec{x})=0.
\]
In conclusion then, with these boundary conditions, the new part of the bulk metric at order $\ep^n$ is therefore
\begin{align}
\label{finalsoln}
g_{r\mu}^\n &= 0, \nn\\
g_{\t\t}^\n &= (1-r/r_c)F_\t^\n(\t,\vec{x})+\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' (\hat{R}_{ii}^\n-r\hat{R}_{rr}^\n-2\hat{R}_{r\t}^\n),\nn\\
g_{\t i}^\n &= (1-r/r_c)F_i^\n(\t,\vec{x})-2\int_r^{r_c}\d r'\int_{r'}^{r_c}\d r'' \hat{R}_{ri}^\n, \nn\\
g_{ij}^\n &= -2\int^{r_c}_r\d r' \frac{1}{r'}\int^{r'}_{0}\d r'' \hat{R}_{ij}^\n,
\end{align}
where the arbitrary functions are
\[
\label{F_relations}
F_\t^\n(\t,\vec{x}) = \beta_2^\n(\t,\vec{x})+\delta P^\n(\t,\vec{x}), \qquad F_i^\n(\t,\vec{x}) = \beta_{4i}^\n(\t,\vec{x})-2\delta v_i^\n(\t,\vec{x}).
\]
We will discuss gauge choices to fix these arbitrary functions in the following section.
As emphasized above,
at each order $\ep^n$ this integration scheme generates a solution that is regular on the future horizon $r=0$.
If $\hat{R}_{\mu\nu}^\n$ is a polynomial in $r$ that is regular at $r=0$, then $g_{\mu\nu}^\n$ will itself be regular at $r=0$. This in turn implies that at the next order $\hat{R}_{\mu\nu}^{(n+1)}$ will be regular as well. Since for the seed solution $\hat{R}_{\mu\nu}^{(3)}$ is regular, this argument ensures regularity of $g_{\mu\nu}^\n$ at $r=0$ for each $n$.
An additional
interesting feature of the hydrodynamic expansion is fact that any vector constructed from $v_i$, $P$ and their derivatives is necessarily of odd order in $\ep$, while any scalar or rank-two tensor constructed from these variables must necessarily be of even order in $\ep$.
This means that at orders $\ep^n$, where $n$ is odd, the Ricci tensor components $\hat{R}_{rr}^\n$, $\hat{R}_{\t\t}^\n$ and $\hat{R}_{ij}^\n$ and the arbitrary function $F_\t^\n(\t,\vec{x})$ all vanish identically, and so the only nonzero metric component generated by the integration scheme \eqref{finalsoln} is $g_{\t i}^\n$. Correspondingly, at orders $\ep^n$ where $n$ is now even, $\hat{R}_{\t i}^\n$ and $F_i^\n(\t,\vec{x})$ are zero, and so and so the only nonzero metric components generated by the integration scheme are $g_{\t\t}^\n$ and $g_{ij}^\n$.
\section{Gauge choices for the fluid}
\label{gauge-choice}
In this section we discuss the gauge choice for the fluid so as fix the arbitrary functions $F_a^\n(\t,\vec{x})$ appearing in the integration scheme \eqref{finalsoln}. In preparation, let us begin by evaluating the contribution $\delta T_{ab}^\n$ to the Brown-York stress tensor arising from the new term $g_{\mu\nu}^\n$ added to the metric at order $\ep^n$.
The variation in the extrinsic curvature of $\Sigma_c$ at order $\ep^n$ due to $g_{\mu\nu}^\n$ is
\[
\delta K^{(n)}_{ab} = \frac{1}{2}\pounds_{N} g_{ab}^\n = \frac{1}{2}N^r\D_r g_{ab}^\n = \frac{1}{2}\sqrt{r_c}\D_r g_{ab}^\n\big|_{\Sigma_c},
\]
where, since $g_{\mu\nu}^\n$ is already of order $\ep^n$, only derivatives with respect to $r$ survive, and the normal is that associated with the background Rindler spacetime.
Thus, we have
\[
\delta K_{\t\t}^\n = -\frac{F_\t^\n(\t,\vec{x})}{2\sqrt{r_c}}, \qquad \delta K_{\t i}^\n = -\frac{F_i^\n(\t,\vec{x})}{2\sqrt{r_c}},
\qquad \delta K^\n_{ij} = +\frac{1}{\sqrt{r_c}}\int_0^{r_c}\d r' \hat{R}_{ij}^\n.
\]
The new term due to $g_{\mu\nu}^\n$ in the Brown-York stress tensor on $\Sigma_c$ at order $\ep^n$ is then
\[
\delta T_{ab}^\n = 2(\g_{ab}\delta K^\n-\delta K_{ab}^\n),
\]
which evaluates to
\begin{align}
\label{deltaT}
& \delta T_{\t\t}^\n = -2\sqrt{r_c}\int_0^{r_c}\d r' \hat{R}_{ij}^\n, \qquad \delta T_{\t i}^\n = \frac{F_i^\n(\t,\vec{x})}{\sqrt{r_c}}, \nn\\[1ex]
&\delta T_{ij}^\n = \frac{F_\t^\n(\t,\vec{x})}{r_c^{3/2}}\,\delta_{ij}+\frac{2}{\sqrt{r_c}}\int_0^{r_c}\d r' (\delta_{ij}\hat{R}_{kk}^\n-\hat{R}_{ij}^\n).
\end{align}
The complete Brown-York stress tensor on $\Sigma_c$ at order $\ep^n$ is then
\[
T^{(n)}_{ab} = \delta T_{ab}^\n + \hat{T}_{ab}^\n,
\]
where $\hat{T}_{ab}^\n$ represents the contribution at order $\ep^n$ due to the metric up to order $\ep^{n-1}$.
\subsection{Fixing \texorpdfstring{$F_i^\n(\t,\vec{x})$}{Fi(t,x)}}
From \eqref{F_relations}, we see that the arbitrary vector function $F_i^\n(\t,\vec{x})$ appearing at odd orders in $\ep$ is related to redefinitions of the fluid velocity. To fix this ambiguity, we define the fluid velocity as the boost from the lab frame to the local rest frame of the fluid, in which the momentum density vanishes
(i.e., $T_{\t i}=0$ in the local rest frame where the fluid velocity is purely timelike).
In an arbitrary frame where the fluid velocity $u^a=\gamma (1, v_i)$, this Landau gauge condition \cite{Landau} then reads
\[
\label{rel_no_mom}
0 = h_a^b T_{bc} u^c, \qquad h^a_b=\delta^a_b + u^a u_b.
\]
At odd orders in $\ep$, the $\t$ component of this equation vanishes identically, since there are no scalars at odd order. The remaining vector component reads
\[
\label{no_momentum}
0 = T_{i\t}^\n+T_{ij}^{(n-1)} v_j + \rho^{(n-1)} v_i,
\]
where $\rho=T_{ab}u^a u^b$ is the energy density in the local rest frame\footnote{While this quantity vanishes for the background solution, it is nonzero at higher order in fluid gradients.}.
Hence at order $\ep^n$ (for odd $n$), we find $F_i^\n(\t,\vec{x})$ is fixed in terms of known quantities according to
\[
\label{Fi_gauge}
0 = \frac{F_i^\n(\t,\vec{x})}{\sqrt{r_c}}+\hat{T}^{(n)}_{i\t}+T^{(n-1)}_{ij}v_j + \rho^{(n-1)}v_i,
\]
where $T^{(n-1)}_{ij}$ and $\rho^{(n-1)}$ denote respectively the full stress tensor and the energy density evaluated at order $\ep^{n-1}$, both of which are known quantities.
If we now move to order $\ep^{n+1}$ (where $n{+}1$ is even), we find the vector component of \eqref{rel_no_mom} vanishes identically (as there are no vectors at even orders), but the $\t$ component is nontrivial and reads
\[
\label{consistency_check}
0 = T^{(n+1)}_{\t\t}+T_{\t i}^{(n)}v_i-r_c\rho^{(n+1)}.
\]
Noting that $\rho = \gamma^2(T_{\t\t}+2T_{\t i}v_i+T_{ij}v_iv_j)$ and that $1-r_c\gamma^2 = -v^2\gamma^2$, one can show that \eqref{consistency_check} is equivalent to
\[
0 = v^2 (T_{\t\t}^{(n-1)}+T_{\t i}^{(n-2)}v_i) + r_c(T_{\t i}^{(n)}+T_{ij}^{(n-1)}v_j)v_i .
\]
Then, assuming \eqref{no_momentum} is satisfied at order $\ep^{n}$, \eqref{consistency_check} reduces to
\[
0 = v^2 (T_{\t\t}^{(n-1)}+T_{\t i}^{(n-2)}v_i - r_c \rho^{(n-1)}).
\]
Thus, \eqref{consistency_check} is satisfied at order $\ep^{n+1}$ provided that it is satisfied at the preceding even order $\ep^{n-1}$. Since the seed solution \eqref{seed} explicitly satisfies \eqref{consistency_check} at order $\ep^2$, we see that the gauge condition \eqref{rel_no_mom} can indeed be consistently imposed at all orders by choosing $F_i^\n(\t,\vec{x})$ according to \eqref{Fi_gauge}.
It is quite nontrivial that we can consistently impose the {\it relativistic} gauge choice \eqref{rel_no_mom} on the dual fluid, since the latter is constructed according to a non-relativistic expansion scheme in which time and spatial derivatives have different order in $\ep$. We will return to this topic in section \ref{Section:Hydro}.
\subsection{Fixing \texorpdfstring{$F_\t^\n(\t,\vec{x})$}{Ft(t,x)}}
\label{Section:Ft}
From \eqref{F_relations}, we see that $F_\t^\n(\t,\vec{x})$, the arbitrary scalar function appearing at all even orders in $\ep$ greater than two, is related to redefinitions of the pressure fluctuation $P$.
To remove this ambiguity, we propose defining the pressure fluctuation $P$ so that the isotropic part of $T_{ij}$ is fixed to be
\[
\label{Ft_gauge}
T^{\mathrm{isotropic}}_{ij} = \(\frac{1}{\sqrt{r_c}}+\frac{P}{r_c^{3/2}}\)\delta_{ij}
\]
to all orders.
From \eqref{deltaT}, it is clear that we can always choose $F_\t^\n(\t,\vec{x})$ at every even order greater than two so as to ensure that $T_{ij}$ receives no higher order corrections that are proportional to $\delta_{ij}$.
While other gauge choices to define the pressure fluctuation $P$ are possible (see footnote \ref{gaug_ch}), the present choice is especially simple from an operational point of view.
\section{Results to \texorpdfstring{$O(\epsilon^6)$}{fifth order}}
\label{explicit}
In this section we now apply the integration scheme developed in the preceding sections to compute the bulk metric and the ensuing fluid stress tensor at orders up to and including $\ep^5$ in four and five bulk spacetime dimensions. We have checked explicitly at the end of the procedure that our solution satisfies the bulk Einstein equations to $O(\ep^6)$ for four and five dimensions.
Since neither the seed metric, nor the algorithm to construct the solution at each step, depends on the dimension; and moreover
scalars, vectors and rank-two tensors formed from $v_i$, $P$ have no special properties in three spatial dimensions
(unlike in two spatial dimensions\footnote{In two spatial dimensions (i.e., $d=2$), $v_i$ has two components and dimension dependent identities for rank 2 tensors exist at order $\epsilon^4$. Notably, $\sigma_{ik}\sigma_{kj}=1/2 \delta_{ij}\sigma_{kl}\sigma_{kl}$ and $\omega_{ik}\omega_{kj}=1/2 \delta_{ij}\omega_{kl}\omega_{lk}$, where $\sigma_{ij}$ and $\omega_{ij}$ are as defined in \eqref{sigmadef}.}),
we expect that the following results are valid in \emph{any} bulk dimension greater than five as well. We have not, however, explicitly checked that the equations of motion hold to $O(\ep^6)$ for bulk dimensions greater than five.
\subsection{Corrections to Navier-Stokes and incompressibility}
As noted in \cite{Bredberg}, imposing conservation of the Brown-York stress tensor at order $\ep^2$ yields the incompressibility condition, while at order $\ep^3$ we recover the Navier-Stokes equation. In section \ref{Section_algorithm}, we saw explicitly how conservation of the Brown-York stress tensor at order $\ep^n$ is required in order to construct the bulk metric at that same order.
Applying the integration scheme described in section \ref{Section_algorithm}, at order $\ep^4$ we obtain the following corrections to the incompressibility equation:
\[
\label{final_I}
\partial_i v_i = \frac{1}{r_c} v_i \partial_i P -v_i \partial^2 v_i +\frac{1}{2}\sigma_{ij}\sigma_{ij} + O(\ep^6),
\]
where the fluid shear $\sigma_{ij}$ and vorticity $\omega_{ij}$ are given by
\[
\label{sigmadef}
\sigma_{ij} = 2\D_{(i}v_{j)} \equiv \D_iv_j+\D_jv_i, \qquad \omega_{ij} = 2\D_{[i}v_{j]} \equiv \D_iv_j-\D_jv_i.
\]
As \eqref{final_I} is a scalar equation, there are no corrections at odd orders, and so the next corrections appear at order $\ep^6$.
As expected, the Navier-Stokes equations are modified at order $\ep^5$:
\begin{align}
\partial_\tau v_i +v_j \partial_j v_i -r_c \partial^2 v_i +\partial_i P &= -\frac{3r_c^2}{2}\partial^4 v_i+2r_cv_k \partial^2 \partial_k v_i+r_c \sigma_{ik}\partial_l \sigma_{kl}-\frac{5r_c}{2}\omega_{ik}\partial_l \sigma_{kl} \nn \\
&\quad -\frac{3r_c}{4}\partial_i (\sigma_{kl}\sigma_{kl}) - \frac{5r_c}{8}\partial_i (\omega_{kl}\omega_{lk})+r_c \sigma_{kl} \partial_k \sigma_{li}-2v_k \partial_k \partial_i P \nn \\
&\quad -2 (\partial_k v_i)\partial_k P-P\partial^2 v_i - \frac{1}{2}v^2 \partial^2 v_i - \frac{1}{2}(\partial_k \sigma_{il})v_k v_l+\frac{1}{2}(\partial_k \omega_{il}) v_k v_l \nn \\
&\quad +2(\partial_k v_i )\omega_{kl}v_l + \frac{1}{r_c}(P+v^2) \partial_i P - \frac{1}{r_c}v_i\partial_\tau P + O(\epsilon^7)\, .\label{final_NS}
\end{align}
Since this is a vector equation, there are no corrections at even orders, hence the next corrections will be at order $\ep^7$.
In writing the form of these correction terms, we eliminated $\D^2 P$ terms using the relationship
\[
\label{derived_eqn}
\D^2 P + (\D_iv_j)(\D_jv_i)=O(\ep^6),
\]
which follows from taking the divergence of Navier-Stokes equation and using incompressibility.
\subsection{Metric at higher orders}
The initial seed metric up to $\ep^2$ is given in \eqref{seed}. Applying our solution algorithm for the bulk geometry, at order $\ep^3$ the only nonvanishing components of the metric are
\[ \label{g3}
g^{(3)}_{\t i} = \frac{(r-r_c)}{2r_c}\Big[(v^2+2P)\frac{2v_i}{r_c}+4\D_iP -(r+r_c)\D^2 v_i\Big].
\]
Note that only the last of these terms was found in \cite{Bredberg}; the first two terms are required in order to impose the fluid gauge condition \eqref{no_momentum} at order $\ep^3$.
At $\ep^4$, the only nonvanishing components of the metric are
\[ \label{g41}
g^{(4)}_{\tau\tau} = -\frac{(r-r_c)^3}{8 r_c^2}(\omega_{kl}\omega_{lk}) +\frac{(r-r_c)^2}{8 r_c}( 8 v_k\partial^2 v_k + \sigma_{kl}\sigma_{kl})-\frac{(r-r_c)}{r_c} F_{\tau}^{(4)},
\]
where
\[
F^{(4)}_{\tau} = \frac{9}{8r_c}v^4+\frac{5}{2r_c}Pv^2+\frac{P^2}{r_c}-2r_c v_i \partial^2 v_i - \frac{r_c}{2}\sigma_{kl}\sigma_{kl}-2\partial_\tau P + 2v_k \partial_k P,
\]
is fixed by the gauge condition. Also,
\begin{align}
g^{(4)}_{ij} = &(1-\frac{r}{r_c}) \Big[ \frac{1}{r_c^2}v_i v_j (v^2+2P)+\frac{2}{r_c}v_{(i}\partial_{j)}P-4\partial_i\partial_j P -\frac{1}{2}\sigma_{ik}\sigma_{kj}+\frac{r-5r_c}{4r_c}\omega_{ik}\omega_{kj} \nonumber\\
&\quad + \sigma_{k(i}\omega_{j)k}-\frac{r+r_c}{r_c}v_{(i}\partial^2 v_{j)}+\frac{r+5r_c}{4}\partial^2\sigma_{ij}-\frac{1}{r_c}v_{(i}\partial_{j)}v^2-\frac{1}{2r_c}\sigma_{ij}(v^2+2P)\Big]. \label{g42}
\end{align}
At next order we find that the only nonvanishing components of the metric are
\begin{align} \label{g5}
g^{(5)}_{\tau i} =(1-\frac{r}{r_c}&) \Big[ -\frac{(r+2r_c)(r+5r_c)}{12}\partial^4 v_i + \frac{r-3r_c}{4}v_k \partial^2 \sigma_{ik}-\frac{1}{2}(r+r_c)v_k \partial^2 \omega_{ik}+r_c \sigma_{ik}\partial_l \sigma_{kl} \nn \\
&\quad - \frac{(r+r_c)(r+4r_c)}{4r_c}\omega_{ik}\partial_l \sigma_{kl}-\frac{5r+r_c}{8}\partial_i (\sigma_{kl} \sigma_{kl})+\frac{r^2-7r_c r-4 r_c^2}{16 r_c}\partial_i (\omega_{kl}\omega_{lk}) \nn \\
&\quad +\frac{r+r_c}{2}\sigma_{kl}\partial_k \sigma_{li} - \frac{r-3r_c}{r_c}v_k \partial_k \partial_i P - \sigma_{ik}\partial_k P + \frac{r}{r_c}\omega_{ik}\partial_k P + \frac{r-2r_c}{r_c}\partial^2 v_i P \nn \\
&\quad + \frac{r}{2r_c}v^2 \partial^2 v_i + \frac{r}{r_c}v_i v_l \partial_k \sigma_{kl} - \frac{r+r_c}{4r_c}\partial_k \sigma_{il}v_k v_l + \frac{r+r_c}{4r_c}\partial_k \omega_{ik}v_k v_l + \frac{1}{2}\sigma_{ik}\sigma_{kl}v_l \nn \\
&\quad + \frac{r+2r_c}{2r_c} \sigma_{ik} \omega_{kl}v_l - \frac{1}{2}\omega_{ik}\sigma_{kl}v_l - \frac{3r-3r_c}{4r_c}\omega_{ik}\omega_{kl}v_l + \frac{r+3r_c}{8r_c}\sigma_{kl}\sigma_{kl}v_i \nn \\
&\quad - \frac{(r-r_c)^2}{8r_c^2}\omega_{kl}\omega_{lk}v_i - \frac{1}{r_c^2}P^2 v_i - \frac{5}{2r_c^2}P v_i v^2 - \frac{9}{8r_c^2}v_i (v^2)^2+\frac{1}{r_c}P \partial_i P \nn \\
&\quad + \frac{r-r_c}{r_c^2}\partial_i v^2 P- \frac{1}{r_c}v_k \partial_k P v_i - \frac{r-2r_c}{r_c^2}v_k \sigma_{ki}P - \frac{1}{r_c}\partial_\tau P v_i + \frac{1}{r_c}\sigma_{ik}v_k v^2 \nn\\
&\quad + \frac{r}{2r_c^2}\omega_{ik}v_k v^2 + \frac{1}{2r_c}v_l \sigma_{kl}v_l v_i
\Big]\, .
\end{align}
Inspecting these results, we see that the bulk metric takes the form of a polynomial in $r$, and as such we would expect it to have only a finite radius of convergence centered about $r=r_c$.
Note however that this location is arbitrary: by applying the scaling $(r,\tau,x_i,v_i,P)\rightarrow (\lambda^2 r, \tau, \lambda x_i, \lambda v_i, \lambda^2 P)$ we may set $r_c$ to any chosen value.
Let us also mention that, while we verified that the $d=4$ solution is Petrov II up to order $\epsilon^{12}$, as noted in \cite{Bredberg}, the metric fails to be of Petrov II type at order $\epsilon^{14}$.
Finally, let us comment on the location of the Rindler horizon for the full solution. This can be
worked out perturbatively in the non-relativistic expansion, as in the analogous computation in \cite{Bhattacharyya:2008xc}.
Let $r=R(\tau,x_i)$ be this position, where $R$ is constructed from $v_i$, $P$ and their derivatives.
The terms without any derivatives follow from the equilibrium solution, which yields $R_{eqm} = r_c [1-(1+P/r_c)^{-2}]$. The remaining gradient terms may then be determined by imposing at each order in $\epsilon$ that the hypersurface be null: $0=g^{\mu\nu}\partial_\mu (r-R)\partial_\nu (r-R)$. To $O(\ep^4)$, there are no gradient terms that can be constructed, hence $R=2P+O(\ep^4)$. To obtain the non-trivial gradient terms appearing at $\epsilon^4$ order, however, would require knowing $g^{rr}$ to $\epsilon^6$ order.
\subsection{Stress tensor at higher orders}
The stress tensor of the seed metric was given in \eqref{seed_stress}. At order $\epsilon^3$, this stress tensor gets corrected by the following expression
\begin{align}
T^{(3)}_{ab}\d x^a \d x^b =2 r_c^{-3/2}\Big[r_c\sigma_{ik}v_k-(v^2+P)v_i\Big]\d x^i\d\t \, .\label{T3}
\end{align}
At order $\epsilon^4$, the only nonvanishing components of the stress tensor that receive a correction are the scalar $T^{(4)}_{\tau \tau}$ and tensor parts $T^{(4)}_{ij}$ as follows,
\begin{align}
T^{(4)}_{ab}\d x^a \d x^b &= r_c^{-3/2}\Big[v^2(v^2+P)-\frac{r_c^2}{2}\sigma_{ij}\sigma_{ij}-r_c \sigma_{ij}v_iv_j\Big]\d\t^2\nn \\[1ex]
&\quad + r_c^{-5/2}\Big[ v_iv_j (v^2+P) +2r_c v_{(i}\partial_{j)}P -4 r_c^2 \partial_i\partial_j P-\frac{r_c^2}{2}\sigma_{ik}\sigma_{kj}-r_c^2 \omega_{ik}\omega_{kj} \nn\\[1ex]
&\quad +r_c^2\sigma_{k(i}\omega_{j)k} -2r_c^2 v_{(i}\partial^2 v_{j)}+\frac{3r_c^3}{2}\partial^2 \sigma_{ij}-r_c v_{(i}\partial_{j)}v^2 -\frac{r_c}{2}\sigma_{ij}v^2 \Big] \d x^i\d x^j \, .\label{T4}
\end{align}
At order $\ep^5$, we obtain
\begin{align}
T^{(5)}_{ab}\d x^a \d x^b &= 2 r_c^{-5/2}\Big[ -\frac{3r_c^3}{2}v_k \partial^2 \sigma_{ik} +4 r_c^2 v_k \partial_k \partial_i P + r_c^2 v^2 \partial^2 v_i
+r_c^2 v_i v_l \partial_k \sigma_{kl} + \frac{r_c^2}{2}\sigma_{ik}\partial_k v^2 \nn \\
& \quad - \frac{r_c^2}{2}\omega_{ik} \sigma_{kl}v_l +r_c^2 \omega_{ik}\omega_{kl}v_l + \frac{r_c^2}{2}\sigma_{kl}\sigma_{kl}v_i - v_i v^2 (P+v^2)-r_c v^2\partial_i P \nn \\
&\quad - r_c v_i v_k \partial_k P +r_c \sigma_{ik}v_k v^2+\frac{r_c}{2}\omega_{ik}v_k v^2 + \frac{r_c}{2}v_k \sigma_{kl}v_l v_i \Big] d\tau dx^i \, .\label{T5}
\end{align}
To simplify the form of these expressions, we have made use of the constraint equations \eqref{final_I} and \eqref{final_NS} below in such a manner that all $\tau$ derivatives of $v_i$ do not appear in the final form of the expressions, as well as equations deriving from these such as \eqref{derived_eqn} in such a manner that $\D^2 P$ does not appear in the final form of the expressions.
\section{Characterising the dual fluid}
\label{Section:Hydro}
Earlier, we noted how the Brown-York stress tensor of the seed metric could be obtained from the $\ep$-expansion of the hydrodynamic stress tensor for a relativistic fluid.
Moreover, when constructing the bulk solution, we found one can consistently choose a gauge such that $u^aT_{ab}h^c_c=0$ to all orders in $\ep$.
Taken together, these observations suggest that it might in fact be possible to recover our full gravitational results for the Brown-York stress tensor from
the $\ep$-expansion of some appropriately chosen relativistic hydrodynamic stress tensor.
Examining the results of the previous section more closely, we find that the energy density in the local rest frame is given by
\[ \label{rho_lead}
\rho = T_{ab}u^au^b = -\frac{1}{2\sqrt{r_c}}\sigma_{ij}\sigma_{ij}+O(\epsilon^6).
\]
In particular, we see that the energy density vanishes for equilibrium configurations in which the fluid velocity is everywhere constant.
In the following section, we discuss the theory of relativistic hydrodynamics at second order in fluid gradients, paying special attention to the modifications necessitated by the vanishing equilibrium energy density.
We will then proceed to match our proposed relativistic hydrodynamic stress tensor with the Brown-York stress tensor derived from our gravitational calculations above, permitting the identification of second-order transport coefficients.
\subsection{Relativistic hydrodynamics for vanishing equilibrium energy density}
Defining the relativistic fluid velocity $u^a$ as in \eqref{rel_no_mom}, so that the momentum density in the local rest frame vanishes,
the fluid stress tensor takes the general form
\[
\label{hydroT}
T_{ab} = \rho u_a u_b + p h_{ab} + \Pi^\perp_{ab}, \qquad u^a\Pi^\perp_{ab}=0.
\]
Here, $h_{ab} = \g_{ab}+u_au_b$ is the induced metric on surfaces orthogonal to the fluid velocity, and $p$ represents the pressure of the fluid in the local rest frame. The term $\Pi^\perp_{ab}$ encodes dissipative corrections and may be expanded in gradients of the fluid velocity.
Since in the present case the equilibrium energy density vanishes, we must also expand the energy density $\rho$ in terms of gradients of the fluid velocity.
Inserting \eqref{hydroT} into the constraint \eqref{Ham_constraint} deriving from the bulk Hamiltonian constraint, we find
\[
\label{rho_result}
0 = \rho \big( (d-1)\rho+2d p+2\Pi^{\perp}\big) + d\Pi^\perp_{ab}\Pi^{\perp ab}-(\Pi^\perp)^2,
\]
where $\Pi^\perp = h^{ab}\Pi^\perp_{ab}$.
This relation then fully determines the energy density $\rho$ in terms of $p$ and $\Pi^\perp_{ab}$ (remembering that we must select the solution branch corresponding to a zero equilibrium energy density). Equation \eqref{rho_result}
therefore plays a role somewhat analogous to that of the equation of state for a conventional fluid.
To write down the gradient expansions for $\rho$ and $\Pi^\perp_{ab}$ precisely, it is useful to first consider the equations of motion at lowest order in fluid gradients: noting that $\rho$ has no component at zeroth order in gradients, these are
\[
\label{zerograd}
0 = u^a\D^b T_{ab} = -p \D_a u^a + O(\D^2), \qquad 0=h_a^b\D^cT_{bc} \quad \Rightarrow\quad D u_a =- D^\perp_a \ln p +O(\D^2),
\]
where we have omitted terms of second and higher order in fluid gradients, and we have defined $D^\perp_a \equiv h_a^b\D_b$ and $D \equiv u^a\D_a$. We may use these equations to simplify the form of the possible coefficients that appear at a given order in the gradient expansion: for example, we see that the only possible first order term in the gradient expansion for $\rho$, namely $\D_a u^a$, in fact vanishes at this order. The expansion for $\rho$ therefore starts at second order.
As shown in appendix \ref{hydro_app}, making use of the relations \eqref{zerograd} we may write down the following complete basis for $\Pi^\perp_{ab}$ up to second order in gradients,
\begin{align}
\label{PiT}
\Pi^\perp_{ab} &= -2\eta \K_{ab} + c_1 \K_a^c\K_{cb} + c_2 \K_{(a}^c\Omega_{|c|b)} + c_3 \Omega_a^{\,\,\,c}\Omega_{cb} +
c_4 h_a^ch_b^d\D_c\D_d \ln p \nn\\
& \quad + c_5 \K_{ab}\,D\ln p + c_6 D^\perp_a \ln p \,D^\perp_b\ln p + O(\D^3),
\end{align}
where $\eta$ is the relativistic kinematic viscosity and the $c_1$, $c_2$, etc., are the corresponding transport coefficients at second order.
Here, we are restricting to flat space\footnote{For curved backgrounds additional terms involving the Riemann tensor appear, see e.g., \cite{Starinets}.}, and we have defined the relativistic shear and vorticity according to
\[
\label{KOdef}
\K_{ab} = h_a^c h_b^d\D_{(c}u_{d)}, \qquad \Omega_{ab} = h_a^c h_b^d\D_{[c}u_{d]},
\]
where symmetrisation and anti-symmetrisation are defined as in \eqref{sigmadef}.
Note also we have not included any terms in \eqref{PiT} proportional to some second-order scalar times $h_{ab}$ itself: this is because our definition of the pressure fluctuation $P$ in \eqref{Ft_gauge} forbids all such terms\footnote{Alternative gauge choices for $P$ are however possible. In particular, one may send $P \rightarrow P - r_c^{3/2}\Pi^\perp/d$, which would render the dissipative part $\Pi^\perp_{ab}$ transverse traceless. This would correspond to the gauge choice \label{gaug_ch} $h^{ab}T_{ab}=d\big(r_c^{-1/2}+r_c^{-3/2}P\big)$.}, as may be seen by going to the rest frame in which $h_{ij}$ reduces to $\delta_{ij}$.
To determine the energy density $\rho$ in the local rest frame, we now simply insert the expansion \eqref{PiT} into the constraint \eqref{rho_result}.
Expanding to second order in fluid gradients, we find
\[
\label{d_result1}
\rho = -\frac{2\eta^2}{p}\K_{ab}\K^{ab}+O(\D^3).
\]
\subsection{Determination of transport coefficients}
\label{Subsection:Hydro_ep}
To test the proposed relativistic hydrodynamic expansion given by \eqref{hydroT}, and to identify the transport coefficients appearing in \eqref{PiT}, we now perform a second expansion of \eqref{PiT} in $\ep$ up to $O(\ep^6$). We will then backsubstitute into \eqref{hydroT}, and compare with our results for the Brown-York stress tensor in \eqref{T3}, \eqref{T4} and \eqref{T5}.
In comparing the two expansions, it is useful to note that the terms associated with the coefficients $c_5$ and $c_6$ both vanish to $O(\ep^6)$, and so are not constrained by the gravitational analysis above. Nevertheless, if required, these transport coefficients may be straightforwardly evaluated by extending the gravitational analysis to the appropriate order.
In the following analysis, we expect to be able to recover all of the terms in the Brown-York stress tensor containing up to two gradients, commensurate with the second-order accuracy of our corresponding hydrodynamic expansion. Note, however, that the number of gradients in a given relativistic term may increase after expanding in $\ep$, since we make free use of the incompressibility and Navier-Stokes equations to simplify our expressions. Crucially though, the order in gradients can never decrease as we go from the relativistic expansion to the $\ep$-expansion.
Examining the terms up to $\ep^2$ order, as noted earlier we find
\[
\label{P_def}
p = \frac{1}{\sqrt{r_c}} + \frac{P}{r_c^{3/2}},
\]
i.e., the pressure $p$ of the dual fluid consists of an equilibrium part $r_c^{-1/2}$ (as expected from the result for Rindler space) and a fluctuation $r_c^{-3/2}P$ at order $\ep^2$.
The nature of our gauge choice \eqref{Ft_gauge} is such that this relation is exact and receives no corrections at higher orders.
The full hydrodynamical stress tensor \eqref{hydroT} expanded to $\ep^2$ order is then
\begin{align}
T^{\mathrm{hydro}}_{ab}\d x^a \d x^b &= \frac{\d \vec{x}^2}{\sqrt{r_c}}
- \frac{2 v_i}{\sqrt{r_c}}\,\d x^i\d \t
+\frac{v^2}{\sqrt{r_c}}\,\d\t^2 + r_c^{-3/2}\Big[P\delta_{ij}+v_iv_j-2\eta r_c \D_i v_j\Big]\d x^i\d x^j +O(\ep^3),
\end{align}
from which we immediately note $\eta=1$, although we will leave $\eta$ explicit in the following formulae.
Evaluating now the terms at $\ep^3$ order, we find
\[
\label{NewT3}
T_{ab}^{(3)\,\mathrm{hydro}}\d x^a \d x^b = 2 r_c^{-3/2}\Big[\eta r_c\sigma_{ik}v_k-(v^2+P)v_i\Big]\d x^i\d\t ,
\]
which is straightforwardly consistent with \eqref{T3} above.
At $\ep^4$ order, we obtain
\begin{align}
T^{(4)\,\mathrm{hydro}}_{ab}\d x^a \d x^b &= r_c^{-3/2}\Big[v^2(v^2+P)-\eta r_c \sigma_{ij}v_iv_j - \frac{r_c^2}{2}\sigma_{ij}\sigma_{ij}
\Big]\d\t^2\nn \\[1ex]
&\quad + r_c^{-5/2}\Big[ v_iv_j (v^2+P) +2\eta r_c v_{(i}\partial_{j)}P + c_4 r_c^{3/2} \partial_i\partial_j P+\frac{c_1}{4}r_c^{3/2}\sigma_{ik}\sigma_{kj}
\nn\\[1ex]
&\quad\qquad + \frac{c_3}{4}r_c^{3/2} \omega_{ik}\omega_{kj}
-\frac{c_2}{4}r_c^{3/2}\sigma_{k(i}\omega_{j)k} -2\eta r_c^2 v_{(i}\partial^2 v_{j)} -\eta r_c v_{(i}\partial_{j)}v^2 \nn\\[1ex]
&\quad\qquad
-\frac{r_c}{2}\eta\sigma_{ij}v^2 \Big] \d x^i\d x^j \, .\label{newT4}
\end{align}
Comparing with Brown-York stress tensor \eqref{T4}, we find exact agreement upon setting
\[
\label{transport_coefficients}
c_1= -2\sqrt{r_c}, \quad c_2=c_3=c_4=-4\sqrt{r_c}.
\]
In fact, the only term in \eqref{T4} not captured by the hydrodynamic expansion \eqref{newT4} is the term proportional to $\partial^2\sigma_{ij}$: as this term is of third order in fluid gradients, however, we do not expect to be able to reproduce it from our second-order hydrodynamic expansion.
Finally, at order $\ep^5$, we obtain
\begin{align}
T^{(5)\,\mathrm{hydro}}_{ab}\d x^a \d x^b &= 2 r_c^{-5/2}\Big[ -c_4r_c^{3/2} v_k \partial_k \partial_i P + \eta r_c^2 v^2 \partial^2 v_i
+\eta r_c^2 v_i v_l \partial_k \sigma_{kl} - \frac{c_2}{8}r_c^{3/2}\sigma_{ik}\partial_k v^2 \nn \\
& \quad +\frac{1}{8}(c_2-2c_1)r_c^{3/2}\sigma_{ij}\sigma_{jk}v_k + \frac{c_2}{8}r_c^{3/2}\omega_{ik} \sigma_{kl}v_l
-\frac{c_3}{4}r_c^{3/2} \omega_{ik}\omega_{kl}v_l \nn\\
&\quad + \frac{r_c^2}{2}\sigma_{kl}\sigma_{kl}v_i
- v_i v^2 (P+v^2)-\eta r_c v^2\partial_i P \nn\\
&\quad
- \eta r_c v_i v_k \partial_k P +\eta r_c \sigma_{ik}v_k v^2+ \frac{r_c}{2}\eta\omega_{ik}v_k v^2 + \frac{r_c}{2}\eta v_k \sigma_{kl}v_l v_i \Big] d\tau dx^i \, .\label{newT5}
\end{align}
Checking this result against \eqref{T5}, we find that we can indeed reproduce all terms (again, apart from the single third-order term proportional to $v_k\partial^2\sigma_{ik}$) with the assignments \eqref{transport_coefficients}.
In conclusion then, we have seen how our seemingly complicated results \eqref{T3}, \eqref{T4} and \eqref{T5} for the Brown-York stress tensor may be recovered from the $\ep$-expansion of the simple relativistic hydrodynamic stress tensor \eqref{hydroT}, where the relevant transport coefficients in \eqref{PiT} are given by \eqref{transport_coefficients}.
\section{Models for the dual fluid}
\label{Section:Models}
In this section, we present
a simple Lagrangian model for the dual fluid.
Since in general one would not expect to be able to reproduce the dissipative part of the stress tensor from such a model (without first coupling to a heat bath),
we focus on reproducing the non-dissipative piece of the stress tensor,
namely
\[
\label{eqT}
T_{ab} = p h_{ab}.
\]
This corresponds to a fluid with nonzero pressure but a vanishing
energy density in the local rest frame, $\rho = T_{ab} u^{a} u^{b} = 0$.
Note that this stress tensor also satisfies the `equation of state'
following from the Hamiltonian constraint, $d T_{ab} T^{ab} = T^2$.
The equations of motion follow from the conservation
of $T_{ab}$, and are then those in
\eqref{zerograd} but with no higher-derivative corrections, i.e.,
\[ \label{feq}
\partial^a u_a =0, \qquad D u_a =- D^\perp_a \ln p,
\]
where $D^\perp_a \equiv h_a^b\D_b$ and $D \equiv u^a\D_a$.
In the previous section, we showed that the stress
tensor of the dual fluid (including dissipative terms) may be obtained from a
non-relativistic limit of the stress tensor for a relativistic fluid.
The fluid velocity spontaneously breaks relativistic invariance, however,
and so we are led to consider a relativistic Lagrangian in
which the relativistic symmetry is spontaneously broken by a background value
for the field.
The following scalar field action satisfies all requirements,
\[
\label{sqrt_action}
S = \int \d^{d+1}x \sqrt{-\g} \sqrt{-(\D\phi)^2}.
\]
The field equations are given by
\begin{equation}
\nabla^a u_a =0, \qquad u_a = \frac{\D_a \phi}{\sqrt{X}}, \label{fluid-vec}
\end{equation}
where $X = -(\D\phi)^2$. Note that $u_a$ satisfies,
\[
u^a u_a = -1.
\]
The stress tensor is given by
\[
T_{ab} = \sqrt{X}\g_{ab}+\frac{1}{\sqrt{X}}\D_a\phi\D_b\phi = \sqrt{X} h_{ab},
\]
This is precisely of the form (\ref{eqT}) with $p = \sqrt{X}$
and consequently $T_{ab}$ indeed satisfies the zero energy
density condition, $\rho = T_{ab} u^{a} u^{b} = 0$, and
the quadratic constraint $d T_{ab} T^{ab} = T^2$. By construction, the
stress tensor is conserved and the equations in (\ref{feq}) are
therefore reproduced.
Note that \eqref{fluid-vec} precisely defines the fluid velocity in terms of a potential
(i.e., the fluid motion corresponds to potential flow).
We also note that taking a generic Lagrangian density $\mathcal{L}(X,\phi)$ and then imposing
\[
0 = T_{ab}u^au^b = \(-2\frac{\delta \mathcal{L}}{\delta \g^{ab}}+\g_{ab}\mathcal{L}\)\frac{\D^a\phi\D^b\phi}{X}
= 2X\frac{\delta\mathcal{L}}{\delta X}-\mathcal{L}
\]
uniquely picks out the square root action \eqref{sqrt_action}, up an
integration constant which can be absorbed into a field redefinition.
The equilibrium configuration with pressure $p=r_c^{-1/2}$
in the rest frame\footnote{The background solution corresponding to the equilibrium configuration
in an arbitrary frame is obtained by Lorentz transforming the r.h.s.~of
(\ref{bkd}). Note that in such a case $T_{\tau \tau} \neq 0$ even though the
energy density is zero.} corresponds to the solution,
\[ \label{bkd}
\phi = \t.
\]
(Recall that the background metric is given by
$\g_{ab}\d x^a\d x^b=-r_c\d\t^2+\d x_i\d x^i$.)
This solution spontaneously breaks Lorentz invariance,
as required.
To model a fluid in its rest frame with small pressure fluctuations about
a constant
equilibrium value of $r_c^{-1/2}$, we must set
\[
\phi = \t + \delta\phi(\t,\vec{x}),
\]
The pressure is then given by
\[
p = \sqrt{X} = \frac{1}{\sqrt{r_c}}(1+2\delta\dot{\phi}+\delta\dot{\phi}^2-r_c\delta\phi_{,i}\delta\phi_{,i})^{1/2} = \frac{1}{\sqrt{r_c}}+\frac{P}{r_c^{3/2}},
\]
where the last equation serves to define the pressure fluctuation $P$ in terms of the field fluctuation $\delta\phi$, in accordance with \eqref{P_def}.
Similarly, from the components of the relativistic fluid velocity $u_\t = -r_c\g=(1+\delta\dot{\phi})/\sqrt{X}$ and
$u_i = \g v_i = \delta\phi_{,i}/\sqrt{X}$, we may solve for the fluid boost velocity $v_i$ in terms of $\delta\phi$, yielding
\[
v_i = -\frac{r_c\delta\phi_{,i}}{(1+\delta\dot{\phi})}.
\]
In summary then, we have shown that the non-dissipative part of the fluid stress tensor \eqref{eqT}, describing a fluid with vanishing
energy density in the local rest frame, leads naturally to the square-root action \eqref{sqrt_action}.
This action is nonlocal in the sense that the expansion around the
background solution involves an infinite number of derivatives.
The square root action has various interesting features.
The action is real for any timelike $(\partial \phi)$ and the Hamiltonian is positive semi-definite since
\[
{\cal H} = X^{-1} (\partial_i \phi)(\partial^i \phi),
\]
so there is no obvious unitarity problem, despite the unconventional nature of the action.
One can generate a wide variety of field theories by adding
extra matter to this scalar field model, provided that one expands about zero background values of these fields.
For example, consider a generic action
\[ \label{gen-mod}
S = \int d^{d+1}x \sqrt{- \gamma} [ f[\psi, A_a, \Phi] \sqrt{- (\partial \phi)^2 + g [\psi,A_a,\Phi]} + h[\psi,A_a,\Phi] ],
\]
with $\psi$ denoting fermions, $A_a$ denoting gauge fields and $\Phi$ denoting scalars. Here the functions
$f[\psi, A_a, \Phi]$, $g [\psi,A_a,\Phi]$ and $h[\psi,A_a,\Phi]$ are only constrained by the requirement that $g$ vanishes
with $f$ finite and $h$ either finite or zero when $\psi = A_a = \Phi = 0$. Choosing a background in which $\phi = \t$ with
all other fields vanishing will always give a zero energy density fluid, as above, for any matter content and choice of functions. Reinstating
the $1/16 \pi G$ in the bulk
will introduce a corresponding prefactor in the action above; as usual, a holographic correspondence
involving classical gravity would be expected to correspond to a dual field theory with a large number of degrees of freedom.
The appearance of the square root in \eqref{sqrt_action} might at first sight suggest a connection with a brane action. Suppose for example that one
considers a $(d+1)$-dimensional brane embedded into a $(d+2)$-dimensional Minkowski target space. The brane embedding would be described by an action
\[
S = - T \int d^{d+1} \xi \sqrt{- \sigma},
\]
where $T$ is the brane tension, $\xi$ are the worldvolume coordinates and $\sigma_{ab} = \D_a Y^{\mu} \D_{b} Y^{\nu} \eta_{\mu \nu}$ is the pulled back
metric, with $Y^{\mu}$ the target space coordinates. Fixing static gauge for the coordinates $(\tau, \vec{x})$ leads to
\[
S = - T \int d^{d+1}x \sqrt{1 + (\D Y)^2 },
\]
where $Y \equiv Y^{d+2}$ is the transverse coordinate to the brane. One can take a tensionless limit of this action in which $T \rightarrow 0$ with
$\varphi = \sqrt{T} Y$ held fixed to eliminate the constant term in the square root
\[
S = - \int d^{d+1}x \sqrt{ (\D \varphi)^2},
\]
but this differs from \eqref{sqrt_action} by the absence of the minus term inside the square root. It follows that the tensionless limit gives rise to a real
action only when $(\D \varphi)^2 > 0$ while the background solution of interest in \eqref{sqrt_action} has $(\D \phi)^2$ timelike, in which case the onshell
brane action would be imaginary. One possibility would be to embed the brane into a flat target space of signature $(d,2)$ so that $Y \rightarrow i Y$
and $\varphi \rightarrow i \varphi \equiv \phi$, reproducing \eqref{sqrt_action} in the tensionless limit.
\
\section{Discussion}
\label{Discussion}
In this paper we established a direct relation between $(d+2)$-dimensional
Ricci-flat metrics and $(d+1)$-dimensional fluids satisfying
the incompressible Navier-Stokes equations, corrected by specific higher-derivative terms.
Our results raise many interesting questions and there are
diverse directions one may wish to further pursue.
Perhaps the most interesting question is whether the correspondence
extends beyond the hydrodynamic regime (on the field theory side)
and/or the classical gravitational description (on the bulk side).
Is there a string embedding of this correspondence?
Even without a string embedding, one may ask how this
correspondence changes if one adds, for example, a bulk stress
tensor or considers higher-derivative corrections to
Einstein gravity.
Will such changes modify the properties of the dual fluid?
The dual fluid has many unconventional features. In particular,
it has zero energy density in equilibrium but nonzero pressure.
What theories have such properties?
One of our most tantalising observations is that there exists
a simple scalar field model that has these properties
and at the same time has no obvious problems with
unitarity, etc. This provides then a candidate model for
the dual to flat spacetime. Can one obtain this model
from branes? Are there any other theories which at and near
equilibrium are described by such a fluid?
This paper has focused on the gravity/fluid
correspondence, computing transport coefficients of the dual fluid
holographically, but it is very exciting to ask how far flat space
holography can be developed. Defining the holographic theory on
$\Sigma_c$, is there a holographic dictionary relating bulk
computations to quantities computable in the dual field theory model?
Here, the construction of the bulk metric was achieved using a
non-relativistic hydrodynamic expansion. We have seen, however,
that there is an underlying relativistic description. Can one
find a manifestly relativistic construction of the bulk metric?
The metric perturbations are by construction regular at the horizon and
throughout the region $r_h < r < r_c$, but, in general, one would expect that the
series expansion only converges in a finite neighbourhood
of the surface $\Sigma_c$. What is the dependence of the radius of convergence
on the defining data on $\Sigma_c$, and what is the range of validity
of the gauge choice for the radial foliation off the surface $\Sigma_c$?
Can one resum the series to obtain a closed-form expression for the
metric?
It is important to note
here that the hydrodynamic expansion can be made around a surface of
{\it arbitrary} radius $r_c$, which in particular can be far from the
horizon. (As $r_{c} \rightarrow \infty$, the pressure and temperature
of the equilibrium fluid tends to zero; the stress tensor of the fluid
is then zero in this limit). The expansion is therefore in general
physically distinct from a near-horizon expansion, although it would
certainly be useful to understand in more detail the limit in which
$r_c$ is taken to be close to the horizon, following \cite{Andy}.
In this paper, we have used flat space in Rindler coordinates as a
seed solution about which the hydrodynamic expansion is made.
As the present construction uses neither the existence of
an event horizon, nor the detailed asymptotic structure of flat space,
there is no reason why it should be applicable only to asymptotically
flat or black hole spacetimes.
In fact, by the equivalence principle,
the construction should hold locally in any small neighbourhood.
Can one patch such a `local' holographic description of neighbourhoods
to obtain a global holographic description of general
spacetimes?
Another generalisation would be to start from a different seed solution.
This should admit at least one Killing vector for the solution to support
equilibrium configurations, but the dual fluid need not live on a flat
spacetime; it could live instead, for example,
on a background such as $R_{\tau} \times S^{d}$. Note that this
would be the geometry of constant-radius slices of Schwarzschild
spacetime.
This work also raises many interesting questions from a general
relativity viewpoint. We have shown explicitly that the data defined
on $\Sigma_c$, together with regularity at the horizon, suffices to
define the solution uniquely as a series expansion in the region
between $\Sigma_c$ and the horizon. Do these solutions exist globally?\footnote{Note that the same question applies also to the standard AdS/fluid
correspondence. In exact parallel, it was shown
in \cite{Bhattacharyya:2008jc}
that the derivative expansion defines the solution uniquely as
a series expansion in the region between the conformal boundary and the
future horizon. It is a non-trivial question whether this provides a global
solution. We thank Bob Wald for discussions regarding this point.}
A priori it would seem
far from obvious that this radial initial value problem is well-posed
and admits a unique solution: it would be instructive to understand
better these issues.
All in all, this correspondence provides an interesting arena
to explore holography for general spacetimes and raises many
interesting questions that we hope to return to in the future.
\acknowledgments
The authors wish to thank Jan de Boer and Michal Heller for discussions.
This work is part of the research program of the `Stichting voor
Fundamenteel Onderzoek der Materie' (FOM), which is financially
supported by the `Nederlandse Organisatie voor Wetenschappelijk
Onderzoek' (NWO). The authors acknowledge support from NWO;
GC and KS via an NWO Vici grant, and PM via an NWO Veni grant.
|
\section{Introduction}
Despite their scarcity, massive stars play an important role
in the history of the Universe. They are the main engines
driving the chemical and dynamical evolution of galaxies, enriching
the interstellar medium with heavy elements, creating H\,{\sc ii}
regions, and exploding as supernovae. In the distant Universe,
they dominate the integrated UV radiation in young galaxies.
Massive stars are possibly key objects for studying and understanding
exciting phenomena such as the cosmic reionisation and
$\gamma$-ray bursters.
In the past decade, our knowledge of the physics of massive stars
has drastically improved because of important developments in
massive star modelling (both in their interiors and outer envelopes),
and continuously higher quality observations collected from
the ground and space. Despite the tremendous progress made so
far,
a number of new challenging issues (regarding both theory and
observations) have emerged that need to be addressed urgently.
Among these are the `weak wind problem' in O-type stars, the wind
momenta of Galactic B-supergiants being significantly lower than
predicted, the behaviour of mass loss at the bi-stability jump, the
effects of rotation and magnetic fields, and wind clumping.
For a comprehensive review on the status-quo of massive star research,
with emphasis on radiatively driven mass loss, the interested reader
is referred to \citep{PVN08}.
For the particular case of O-stars in our Galaxy, the investigation
of the aforementioned open problems, especially those related to stellar
winds, requires a large
sample of objects to be analysed to diminish the error caused by
the uncertain distances. Since Galactic O-type stars constitute
an important reference frame to investigate metallicity effects,
this point is also crucial for extragalactic surveys. Though
easy accessible with small and medium size telescopes, Galactic
O-stars have not been widely observed to derive consisitent stellar
and wind parameters from quantitative spectroscopy, and no
more than 50 objects have been studied in detail
\citep{Herrero02, repo, bouret05, martins05b, dufton06, marcolino}.
\begin{table*}
\begin{center}
\caption[]{Sample stars along with spectral classification.
Low-resolution $quantitative$ and $morphological$ classification
from Mathys \citep{mathys88} and from Walborn \citep{walborn72,
walborn73, walborn02} and Garrison, Hiltner and Schild (1977, GHS),
respectively; high-resolution
$quantitative$ classification from the present study together
with $morphological$ classification based on the original data
($R$=48\,000) and degraded ones ($R$=4000).
Additional
information about $v \sin i$\,\, (in \kms; macroturbulence not accounted
for) and binarity status of the stars is also provided. Binarity
status from speckle interferometry \citep{mason09} and spectroscopy
(previous and present). References for $v \sin i$\,\,- estimates: H
= \citet{howarth97}; M = Markova et al. 2011 (Paper II); P =
\citet{penny96}; U = \citet{UF}; CE = \citet{CE77}. See text for
further information.}
\label{photom}
\tabcolsep1.8mm
\begin{tabular}{lllllrllll}
\hline
\hline
\multicolumn{1}{c}{Star}
&\multicolumn{4}{l}{Previous}
&\multicolumn{1}{c}{$v \sin i$\,}
&\multicolumn{4}{l}{Present study}
\\
\cline{2-5} \cline{7-9}
\multicolumn{1}{c}{}
&\multicolumn{1}{l}{Walborn}
&\multicolumn{1}{l}{Mathys}
&\multicolumn{1}{l}{GHS}
&\multicolumn{1}{c}{Intfr.$^{a}$/Spst.$^{b}$}
&\multicolumn{1}{c}{}
&\multicolumn{1}{l}{Morph.($R_{48000}$)}
&\multicolumn{1}{l}{Quant.}
&\multicolumn{1}{l}{Morph.($R_{4000}$)}
&\multicolumn{1}{c}{Spect.$^{b}$}
\\
\hline
HD~64568 &O3 V((f*)) & &&VS &100, M &O3 V((f*)) & &O3 V((f*)) &SS \\
HD~93204 &O5 V((f)) &O5.5 V((f)) &&VS &115, M &O6 V((f)) &O5.5 V &O6 V((f)) &SS \\
HD~93843 &O5 III(f) &O5.5 III(f) &&VS &100, P &O6 III(fc) &O5.5 III &O5.5 III(fc)&SS \\
CPD~$-$59\,2600&O6 V((f)) & &&VS &142, H &O6 V((f)) &O6.5 V/III&O6 V((f)) &SS \\
HD~63005 & &&O6 V((f)) & & 74, H &O7.5 V((f)) &O6.5 V &O7 V((f)) &SS \\
HD~152723 &O6.5 III(f) &O7 III(f) &&VD, SB1? &123, P &O6.5 III(f) &O6.5 III &O6.5 III(f) &SB1\\
HD~93160 &O6 III(f) & &&VS, SB1? &205, U &O7 III(f) &O7 V &O7 III(f) &SB1\\
HD~94963 &O6.5 III(f) & &&VS &90, CE &O7.5 II(f) &O7-7.5 I/III &O7.5 III(f) &SS\\
CPD\,$-$58\,2620 &O6.5 V((f))& &&VS &60, M &O8 V((f)) &O7 III &O7.5 V((f)) &SS \\
HD~69464 &O6.5 Ib(f) & &&VS &82, P &O7.5 II(f) &O7-7.5 III &O7 III(f) &SB2?\\
HD~93222 &O7 III(f) & &&VS &77, P &O8 III (f) &O7 III &O7.5 III(f) &SS\\
HD~91824 &O7 V((f)) &O7 V ((f)) &&VS &65, H &O8 V ((f)) &O6.5 V &O7 V ((f)) &SS\\
CD\,$-$43\,4690& &O7.5 III(f) && &120, M &O7 III(f) &O6.5 V/III &O7 III(f) &SS \\
HD~92504 &O8.5 V((n)) &O9 V && &200, U &O8.5 V &O9 III/V &O8.5 V &SS \\
HD~151003 &O9 II &O9.5 III &&VS, SB1? &120, U &O9 II &O9 III &O9 II &SB1 \\.
HD~152247 &O9.5 II/III &O9.5 III &&VS, SB2 &120, P &O9 II &O9 III &O9 II &SB2 \\
HD~302505 & &&O8.5~III & & 80, M &O9 III &O8 III &O9 III &SS \\
CPD\,$-$44\,4865& &O9.7 III & & & 80, M &B0 III &O9.5 III &B0 III &SS \\
HD~69106 & & &B0.5 IVnn & &325, H &B0.2 V &O9.7 &B0.2 V &SS \\
\hline
\end{tabular}
\end{center}
$^{a}$ Results from the speckle interferometric survey
of \citet{mason09}: VD = visually double; VS = visually single object.\\
$^{b}$ Spectroscopic binarity status: SS = spectroscopically single
object; SB1/SB2 = single-/double-lined spectroscopic
binary; SB1/2? = possible spectroscopic binary.\\
\end{table*}
The main goal of our project is to increase the number of Galactic
O-type stars with reliably determined physical parameters, using
high-quality spectra and applying the methods of quantitative
spectroscopic analyses. On the basis of new data and incorporating
similar data from previous investigations, we plan to reinvestigate
the latest calibrations for these stars \citep{martins05a}, to address
the important question of macroturbulence and its effect on the
derived $v \sin i$\,, and to investigate the weak wind problem
in stars of solar metallicity. Since the {\it absolute} calibration
of the various relationships between physical parameters and observed
spectral characteristics (e.g., line strengths or, equivalently,
spectral types) requires reliable estimates of all involved quantities,
a thorough investigation of the accuracy of the spectral types and luminosity
classes initially assigned to our targets and an update of their
binary/multiplicity status are required.
In this first paper of the series, the main results from the spectral
classification of the low luminosity stars in our sample (dwarfs and
giants) are presented. A second paper providing the physical properties
of these stars derived by means of the latest version (V10.1) of
the state-of-the-art model atmosphere code {\sc fastwind} \citep{puls05}
is currently in preparation and will be published soon (Markova et al.
2011, hereafter Paper II). A third paper concentrating on the spectral
classification and model atmosphere analysis of the supergiants in
our sample is planned for the near future.
The present paper is structured as follows. In Section~\ref{obs},
we describe the stellar sample and the observational material
underlying the project. In Section~\ref{spec_class}, we outline
the main steps in our classification procedure. Section~ \ref{binary}
deals with the possible binarity/multiplicity among
the sample stars. In
Section~\ref{results}, the correspondence between our
high-resolution $morphological$ and $quantitative$ classification and
the classifications attributed by Walborn \citep{walborn72,
walborn73} and \citet{mathys88} for stars in common is investigated,
and a possible explanation of the derived discrepancies is provided.
The spectral types determined in the framework of the Walborn and
Conti-Mathys schemes from our high-resolution observations are compared
in Section~\ref{quant-morph}. Finally, in Section~\ref{conclusions},
we summarise the main results of our study and comment on some
implications these findings might have in the future. Extensive
comments on each sample star and an atlas of the corresponding
high resolution spectra
are presented in the appendix.
\section{Observations and data reduction}
\label{obs}
The total sample consists of 40 bright ($V$~$\le$~10~mag) O-type
stars in the Milky Way originally classified as dwarfs, giants,
and supergiants, with spectral types ranging from O3 to B0.5.
The targets
were observed with the FEROS spectrograph \citep{kaufer99} at the
ESO/MPG 2.2\,m telescope in La Silla on the 4th and 5th of February,
2004. Each spectrum covers a wavelength range from about 350 to
about 920\,nm with a spectral resolution $R$ = 48\,000. Exposure
times ranging from 400 to 2500\,s were used to obtain a typical
signal-to-noise ratio (hereafter S/N) of 150-200 per resolution
element. One-dimensional, wavelength-calibrated spectra were
extracted using the FEROS pipeline.
For our present study, we considered only the subset of the
sample dwarfs and giants (19 in total) to diminish the
effects of stronger winds on the outcome of the classification
analysis. The stellar IDs along with our present and previous
spectral classifications by Walborn \citep{walborn72, walborn73,
walborn02} and \citet{mathys88} are presented in Table~\ref{photom}.
We note that for three sample stars that have been classified by
neither Walborn nor Mathys, previous classifications based on
low-resolution photographic spectra and MK standards were adopted
(Column 4 of Table~\ref{photom}).
\section{General comments}
\subsection{Spectral classification}
\label{spec_class}
The modern-era O-star spectral classification relies on either
the Walborn or the Conti-Mathys schemes, both being somewhat
related to the MK system defined essentially in terms of standard
stars. In particular, based on a careful investigation of MK
standards observed at a spectral resolution twice that of the
MK atlas, Walborn (1971, 1972, 1973) developed a two-dimensional
empirical system for O-type stars of solar metallicity. This system
was additionally worked out by Walborn and coworkers, and transferred
to digital data of similar \citep{wf90} or higher \citep{walborn02}
resolution. Since the Walborn approach relies on a visual inspection
of the spectra, and the visibility of certain lines and certain
line ratios are used as criteria, it is usually referred to as a
$morphological$ classification. Independently of and simultaneously
with the morphological approach, Conti and collaborators \citep{conti71,
CL74, CF77} developed an alternative method based on logarithmic ratios
of equivalent widths (EW) of certain helium and metal lines, calibrated
against MK spectral types. The $quantitative$ classification system
of Conti was refined by Mathys \citep{mathys88, mathys89}.
To classify our targets, we applied both the Walborn and the
Conti-Mathys systems. In the former case, we considered the visual
impression of our high-resolution spectra and, following the
premises of \citet{wf90}, performed a $morphological$ classification.
However, since the line visibility depends directly on the
data quality, we refrained from using criteria related to
the initial appearance of certain lines and concentrated only on
{\it eye-estimated} line ratios, to avoid possible systematic
effects caused by the higher resolution (and S/N) of our spectra.
We primarily used the He~I to He~II line ratios (particularly
those of both He~I+II~$\lambda$4026\,\, to He~II~$\lambda$4200\,\, and He~I~$\lambda$4471\,\, to He~II~$\lambda$4541\,) to determine
the spectral subtype, except for the hottest stars where criteria
based on the nitrogen ionization equilibrium \citep{walborn02} were
instead applied. For the luminosity class, we followed
the Walborn approach and used different criteria at different
subtypes. For example, at earlier types we considered the effect
of luminosity on the strength of both the N~IV~$\lambda$4058 and He~II~$\lambda$4686\,\,
emission, and the N~V~$\lambda\lambda$4604, 4620 absorption.
At spectral types from O6 to O8, the main luminosity indicators
are He~II~$\lambda$4686\,\, and N~III $\lambda\lambda$4634-4640-4642\,: in dwarfs, the former line appears strongly
in absorption, accompanied by weak N~III emission - the V((f)) category;
in giants, the He~II~$\lambda$4686\,\, absorption weakens and can even vanish while
the N~III $\lambda\lambda$4634-4640-4642\,\, emission strengthens - the III(f) category. At types
O9-B0, the primary luminosity criterion is the strength of the Si~IV doublet
around H$_\delta$\,, whose absorption increases with luminosity.
Our $quantitative$ classification has been performed by exploiting
the logarithmic ratios
$\log W'$ = $\log EW$(HeI$\lambda$4471) $-$ $\log EW$(HeII$\lambda$4542),
which un\-am\-biguously determines the spectral type, and
$\log W''$ = $\log EW$(SiIV$\lambda$4089) $-$ $\log EW$(HeI$\lambda$4143),
which is the main luminosity indicator for stars of spectral types
O7 and later \citep{conti71}. For stars earlier than O7, on the
other hand, the negative luminosity effect in He~II~$\lambda$4686\, with a
demarcation line between class V and III set at
$\log~EW$(HeII4686) =$-$0.25\footnote{ $\log~EW$(HeII4686) larger
than -0.25 determines the luminosity class V.} was used
(\citealt{mathys88} and references therein).
The equivalent widths of the lines involved in the $quantitative$
classification are provided in the appendix (Table~\ref{EW}).
The accuracy of these estimates ranges from 0.02 to 0.07~dex,
with lower values being typical of stronger lines. The
errors in $\log~W'$ and $\log~W''$ are between 0.05 and 0.07~dex
(i.e., less than half a subtype), and about 0.1~dex, respectively.
Our final $morphological$ and $quantitative$ classifications
are listed in Cols. 7 and 8 of Table~\ref{photom}.
Column 6 gives $v \sin i$\,-estimates as adopted
in the present study, primarily drawn from \citet{penny96} and
\citet{howarth97}\footnote{With respect
to our targets, the consistency between the two datasets is
good, to within $\pm$10~\kms.}, if present. Otherwise, individual data
provided by \citet{UF} or derived in Paper II were used instead.
\subsection{Spectroscopic binarity}
\label{binary}
It is well known that spectroscopic binarity can significantly
bias the classification of stars, modifying the visual appearance of
the spectrum and altering the observed equivalent widths of lines
used as spectral type and luminosity class diagnostics. This point is
particularly important in the case of hot massive stars where the
number of double/multiple systems seems to be very large (e.g.,
\citealt{mason09} and references therein).
To investigate the binary status of our targets, we checked
the literature for known/suspected binaries among our sample
stars and found that three of them -- HD~93160, HD~151003
\citep{gies87}, and HD~152723 \citep{full96} -- have been
$suspected$ to be single-lined spectroscopic binaries (SB1),
because of their $V_{\rm r}$\,-variability. Large variations in $V_{\rm r}$\,\
were also detected for another star, HD~64568 \citep{SN86}.
Since $V_{\rm r}$\, - variability can be caused by binary motion, but
may also originate from stellar pulsations and/or wind effects,
we decided to keep the $V_{\rm r}$\,-variables in our target list for
further clarification. Meanwhile, we rechecked the literature
for newer results and found that one of our sample stars
(HD~152247) had been recognized as a double-lined spectroscopic
binary \citep{sana08}; another one (HD~152723) was resolved
as a close double system via $V$-band speckle interferometry,
while for other 12 objects (flagged with ``VS'' in column 5
of Table~\ref{photom} ) a ``null companion detection'' in the
angular separation range 0."035~$\le\rho\le$1."5 and $\Delta m \le$ 3~mag
was reported (\citealt{mason09} and references therein). We kept the
two binaries in the present sample with the primary goal of
investigating the effect of binarity on the outcome of our spectral
classification.
We also tried to constrain the spectroscopic binarity status
of all sample stars by our own. To this end, we adopted the following
criteria as indications for possible binarity: (i) composite spectrum;
(ii) discrepant shifts, widths and/or strengths of lines in the spectrum;
(iii) large differences in $V_{\rm r}$\,\, and $v \sin i$\,\, when measured at different
observational epochs; (iv) periodic variations in photometry and/or radial
velocity; and (v) discrepant spectral classification from various observational
epochs. Since physical phenomena that differ from binarity
can give rise to periodic photometric and $V_{\rm r}$\,-variability while discrepant
spectral classifications might be due to third parameter effects (see below),
these last two criteria were considered only as suggestive.
A comprehensive description of our findings for each star is presented
in the appendix. The spectroscopic status of the targets is listed in
Column 10 of Table~\ref{photom}: ``SS'' denote spectroscopically single
objects, ``SB2'' are double-lined, and ``SB1'' single-lined spectroscopic
binaries.
In summary, following our approach we confirm the binary status
of HD~152723 and HD~152247, provide strong evidence of a companion in
HD~93160 and HD~151003, and suspect duplicity in HD~69464.
\section{Comparison with previous classifications from
low resolution spectroscopy}
\label{results}
\subsection{The present high-resolution morphological
classification and the Walborn results}
\label{firstcomp}
To investigate whether the results derived by means of the
Walborn scheme might be influenced by the spectral resolution,
we
\begin{figure}
\center
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg1a.ps}}
\\
\vspace{-0.5cm}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg1b.ps}}
\caption{{\it Top}: Comparison of our high-resolution {\it
morphological} spectral types and those attributed by Walborn by
means of low-resolution $photographic$ spectra ($\Delta\lambda$ = 1.2 \AA),
for 14 stars (dwarfs and giants) in common. The X-scale corresponds
to spectral types from O2 to B2. Stars considered/suspected to be
SBs are additionally marked with large circles. {\it Bottom}: The
spectral type discrepancy as a function of $v \sin i$\,. To eliminate
binarity effects on the observed $v \sin i$\,, stars considered/suspected
to be SBs have been discarded. Note that the adopted $v \sin i$\,-estimates
do not account for the effects of macroturbulence.
}
\label{SPT_morph}
\end{figure}
compared our $morphological$ spectral types with those attributed
by Walborn, based on low-resolution ($\Delta\lambda$ = 1.2 \AA)
$photographic$ spectra, for 14 stars in common (see Column 2 of
Table~\ref{photom}). As illustrated in Fig.~\ref{SPT_morph}
(top panel), the high-resolution $digital$ classification
tends to result in somewhat later (by up to 1.5 subtypes)
spectral types than those derived by Walborn. This discrepancy
is larger than the error in the $morphological$ classification
(typically less than one subtype), thus significant.
An interpretation in terms of variable wind effects or possible
binarity does not seem likely, since all targets are low luminosity
objects, i.e., have low-density winds, and the majority of them
do not appear to have a companion. In addition, the
distribution of the data suggests that the derived discrepancy might
be present at intermediate subtypes only. Since the main spectral type
indicator at these subtypes is the {\it eye-estimated} ratio of either
He~I+II~$\lambda$4026\,\, to He~II~$\lambda$4200\,, or He~I~$\lambda$4471\,\, to He~II~$\lambda$4541\,, this result implies in
turn
that within the high-resolution digital spectra either He~I+II~$\lambda$4026\,\, and
He~I~$\lambda$4471\,\, appear to be $visually$ stronger or He~II~$\lambda$4200\,\, and
He~II~$\lambda$4541\,\, appear to be $visually$ weaker (or both), compared to their
corresponding strengths as estimated from the $photographic$ spectra.
Apart from stars with discrepant spectral types, our analysis
shows that there are also others with identical $morphological$
classifications as derived from high- (digital) and low-resolution
(photographic) spectra. One of these (HD~64568) is a very early
O3 dwarf (see below); two other stars (HD~152723, O6.5 and
HD~151003, O9) are considered to be spectroscopic binaries
(present study); and the final two are among the fastest rotators
in our sample (CPD -59~2600, O6, $v \sin i$\,=142~\kms and HD~92504, O8.5,
$v \sin i$\,=200~\kms). The latter finding is quite intriguing since it
might imply that the encountered discrepancy is also related to
rotation. We test this possibility in Fig.~\ref{SPT_morph} (lower panel),
where the differences in spectral type are plotted against $v \sin i$\,.
At least for our subsample, the spectral type discrepancy indeed
refers to stars with relatively slow rotation ($v \sin i$\,$\le$120~\kms).
\footnote{Similar results apply for the three stars classified by
\citet{GHS77}, see Table~\ref{photom}, Column 4.} The only exception
is HD~64568 with $v \sin i$\, = 100~\kms, which, however, has not been
classified in terms of the {\it eye-estimated} ratios of He~I to He~II but
instead exploits the relative strengths of nitrogen lines of different
ionisation.
Summarizing,
our analysis shows that using high resolution
$digital$ spectra of O-type stars of solar metallicity results in
somewhat later $morphological$ subtypes, where this discrepancy is
more pronounced in narrow-lined stars.
\begin{figure}
\center
\resizebox{\hsize}{!}
{\includegraphics[width=8.5cm,height=5cm]{15956fg2a.ps}}\\
\resizebox{\hsize}{!}
{\includegraphics[width=8.5cm,height=5cm]{15956fg2b.ps}}
\caption{Examples illustrating the sensitivity of the central
depth of He~I~$\lambda$4471\,\, and He~II~$\lambda$4541\,\, to stellar rotation ($v \sin i$\,=60
(solid), 100 (dotted), 150 (dashed), 200 (dash-dotted) \kms), and
spectral resolution ($R$ = 30\,000 (top),
4\,000 (bottom)). Model profiles from {\sc fastwind}, calculated
at $T_{\rm eff}$\,=38~kK, $\log g$\,=3.7 and log~$Q$=-7.72, with
log~$W'$=-0.19 corresponding to subtype O6.5.
}
\label{res_rot}
\end{figure}
\begin{figure*}
\center
{\includegraphics[width=18cm,height=7cm]{15956fg3.ps}}
\caption{A set of FEROS spectra illustrating the $combined$ effect
of rotation and high spectral resolution on the central depth of
the strategic He~I $\lambda$4026, 4471 and He~II $\lambda$4200, 4541
lines, for stars of similar spectral type (as attributed by
Walborn) and similar $log W'$, as measured from our high-resolution
data. Note that in the Walborn scheme the O7 subtype is defined by
He~I~$\lambda$4471\,~$\approx$~He~II~$\lambda$4541\,, whereas in our spectra He~I~$\lambda$4471\,~ is always deeper
than He~II~$\lambda$4541\,\ (although its EW is actually smaller, log~$W'\le$ 0),
indicating a spectral type of O8-8.5.}
\label{spectra}
\end{figure*}
\paragraph{High-resolution against low-resolution} $morphological$
luminosity classes. A comparison of the luminosity classes
determined in this work with those attributed by Walborn for
14 stars in
common shows perfect agreement for all but two stars -
HD~94963 and HD~69464. This
finding suggests that at least for dwarfs and giants the luminosity
class criteria within the Walborn scheme are not affected by resolution,
rotation, and/or origin of data.
\subsubsection{Influence of rotation and spectral resolution}
\label{rot_effect}
{\it That} rotation should influence the depths of He~I and He~II
lines differently is to be expected from the different intrinsic
widths of these lines, caused by the Stark effect (linear in
He~II and quadratic in He~I). This point is illustrated in
Fig.~\ref{res_rot}, where synthetic profiles for He~I~$\lambda$4471\,\,
and He~II~$\lambda$4541\,, as calculated from a {\sc fastwind} model at $T_{\rm eff}$\,=38~kK,
$\log g$\,=3.7, and log~$Q$=$-$7.72\footnote{$Q$=$\dot M$\,/$R_\star$\,$^{1.5}$
is the optical depth invariant for recombination-based diagnostics
such as H$_\alpha$\,, introduced by \citet{puls96} to characterise the
strength of the wind.
log~$Q$=$-$7.72 ($\dot M$\, in $M_\odot$\,/yr, $R_\star$\, in $R_\odot$\,) represents
the case of a low density wind, originating in a low luminosity
O-type star.}, have been convolved with four values of $v \sin i$\,,
while keeping the spectral resolution fixed at $R$=30\,000 (top panel)
and 4\,000 (bottom panel). The main result of our calculations
is that at the same temperature, surface gravity, chemical
abundances, mass loss rate, and spectral resolving power the
{\it eye-estimated} ratio of He~I~$\lambda$4471\, to He~II~$\lambda$4200\, can indicate different spectral
types, depending on $v \sin i$\,. A comparison of the profiles shown in
Fig.~\ref{res_rot} indicates that the $morphological$ spectral
types for fast ($v \sin i$\,$\ge$150~\kms) and slowly ($v \sin i$\,$\le$100~\kms)
rotating stars can differ by about half (at $R$=4\,000) to about
one and a half (at $R$=30\,000) subtypes, where slowly rotating
stars display a later subtype.
Similar results have been obtained for He~I+II~$\lambda$4026\,\, and He~II~$\lambda$4200\,\,
determining the $morphological$ spectral types from O5 to O6.5.
Illustrative examples of the $combined$ effect of rotation and high
spectral resolution on the visual strength of strategic He~I and
He~II lines are shown in Fig.~\ref{spectra}, where the FEROS spectra
of three narrow-lined sample stars, originally classified as
O6.5 - O7 by Walborn, are displayed along with log~$W'$-values from
our high-resolution EW measurements. The central depths
of He~I~$\lambda$4471\,\, and He~II~$\lambda$4541\,\, lines obviously react in the way predicted by
our model calculations, and indicate a somewhat later spectral
type, namely O8 to 8.5 depending on $v \sin i$\,.
Thus, we conclude that the established discrepancy between
high- and low-resolution $morphological$ spectral types
(mostly for stars of intermediate subtypes) is likely related
(at least in part, see below) to the interplay between
stellar rotation and spectral resolution leading to different
depths of He~I and He~II lines, where the profiles become
deeper at lower rotational speeds and higher resolution. We
emphasize that for stars with identical intrinsic
parameters but different rotational speeds, there will be a
decisive difference in spectral types derived by either
morphological or quantitative classification: whereas the
morphological types will vary according to rotational (and
resolution) effects, types derived by a quantitative
classification will be rather unique, since
rotational broadening preserves the EW and thus $W'$.
\subsubsection{The modern-era $morphological$ classification.}
\label{spectra_degrading}
It is well known that the overall appearance of a stellar
spectrum is strongly dependent on its resolving power. Because
of this, the modern-era morphological classification is
usually performed using (original) spectra that have been
degraded to the resolution of the standards. In this
process, one expects the morphology of the degraded spectra to
closely resemble that of the low resolution classification standards.
To our knowledge, however, this expectation has not been
so far checked using real data.
For a first test of this issue, the original, high-resolution
spectra of our targets were degraded to $R$=4\,000
\footnote{which matches the quality of the photographic spectra
used by Walborn, namely $\Delta\lambda$ = 1.2 \AA.} (using the
IDL procedure ``rebin.pro''), and subsequently classified following
the Walborn criteria. In Fig.~\ref{rbn}, we confront the spectral
types obtained in this way to those attributed by Walborn for stars in
common. Although the agreement between the two data sets has
improved (compared to the one shown in the top panel of
Fig.~\ref{SPT_morph}), the spectral types based on the $degraded$
spectra (Column 9 of Table~\ref{photom}) still tend to be somewhat
later than those derived by means
of the $photographic$ classification, where again this discrepancy
is more significant in narrow-lined stars. To check whether this
finding depends on the procedure used to degrade the spectra, we
repeated the process using an own IDL procedure based on convolution
by Fourier-transforms, and obtained similar results.
\begin{figure}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg4.ps}}
\caption{Comparison of our $morphological$ spectral types,
based on the high-resolution spectra degraded to $R$=4\,000,
and those attributed by Walborn using low-resolution photographic
spectra ($\Delta\lambda$~=~1.2 A) for 14 stars in common. The X-scale
corresponds to spectral types from O2 to B2. Stars considered/suspected
to be SBs are marked with large circles.}
\label{rbn}
\end{figure}
\begin{figure*}
{\includegraphics[width=18cm,height=5cm]{15956fg5.ps}}
\caption{The spectrum of the narrow-lined star CPD-58\,2620
($v \sin i$\,=60~\kms) at the original resolving power $R$=48\,000
and degraded to $R$=3\,000. Spectral classification from
Walborn based on low-quality photographic spectra. Note that
He~I~$\lambda$4471\,\ appears of similar strength as He~II~$\lambda$4541\, in the degraded
spectrum, indicating the O7 subtype.}
\label{resol_effect_2}
\end{figure*}
The effect of significantly different spectral resolving
power on the overall appearance of the spectrum is
illustrated in Fig.~\ref{resol_effect_2}, where
the original, high-resolution spectrum of one sample star
with relatively narrow lines (CPD -58\,2620, $v \sin i$\,=60~\kms),
classified as O6.5 by \citet{walborn73}, is shown together with
its rebinned spectrum\footnote{Here the original spectrum
has been degraded to $R$=3\,000 ($\Delta\lambda$ = 1.5 \AA)
to match the quality of the standards shown in the Walborn and
Fitzpatrick atlas.}. Even at this low resolution,
the spectrum of CPD -58\,2620 obviously does not resemble that of HD~93146,
the classification standard of O6.5V, but is similar to that
of
15~Mon, the classification standard for O7. Thus, and at least
for the case of narrow-lined stars of intermediate spectral type,
it appears that degrading the high-resolution spectra to
the resolution of the Walborn standards does not necessarily
guarantee that they will resemble the morphology of the corresponding
standards, since in this case the comparison objects (the one under
consideration and the standard) might have different $v \sin i$\,. The
larger the difference, the larger the offset in derived spectral
type.
The results outlined above are somewhat surprising since
(i) we degraded our spectra to the resolution of
the photographic data, and (ii) we classified the same stars
as Walborn, i.e., the
resolution and rotational effects should be the same in both
data sets.\footnote{Note that the corresponding spectral types
might still be biased against the standards due to rotational
effects, but both (ours and Walborn's) in the same way, see below.}
Since there are still differences, additional effects must be
at work. What might these effects be?
\begin{itemize}
\item[i)] The remaining discrepancies shown
in Fig.~\ref{rbn} might be caused by the intrinsic differences
in the classification techniques applied in the
$photographic$ and $digital$ classifications: while in
the former case the eye responds to the total flux
transmitted/blocked by the investigated lines (recalling, that
in the early seventies of the last century the classification
was performed by viewing the spectrographs through a microscope),
in the latter it is guided by the central depths of the lines. Consequently,
both estimates might deviate from each other to some degree,
particularly at intermediate subtypes where the differences in
line strength are smaller than in other regions.
\item[ii)] Since O-stars are often embedded in emitting gas,
underestimated nebula contamination of He~I lines in the photographic
spectra might lead to systematically different classifications
from those originating from the high-resolution digital
data.
\item[iii)] Finally, there is a certain chance of line profile
variability. Owing to the systematic character of the discrepancy,
we consider this possibility, however, less likely.
\end{itemize}
Thus, we have to conclude that part of the discrepancies
between
$morphological$ classifications attributed using high-resolution
digital and low-resolution photographic spectra (Fig.~\ref{SPT_morph})
is due to some additional effects to resolution and
rotation alone. This point was also discussed by \citet{walborn10b}.
\subsection{The present high-resolution quantitative
classification and the Mathys results}
\label{mathys}
Since the Conti spectral type classification relies on the
logarithmic ratio of the measured equivalent widths of He~I~$\lambda$4471\,\,
and He~II~$\lambda$4541\,, it is essential to investigate the correspondence
between the present equivalent width scale and that from
similar studies using low-resolution spectra (e.g., \citealt{conti71};
\citealt{CF77}; \citealt{mathys88, mathys89}), to
provide a useful comparison. We limit ourselves to the
study by \citet{mathys88}, which has the largest overlap
with our stellar sample.
In the top panel of Fig.~\ref{comp_EW}, our EW measurements
for He~I $\lambda$4471 (asterisks) and He~II~$\lambda$4541\, (diamonds) are
compared to those from \citet{mathys88} for nine stars in common.
Despite the small number of objects, a clear trend for He~II~$\lambda$4541\,
is visible, indicating that our measurements result in
systematically stronger EWs than derived by Mathys. A linear fit
to the corresponding data confirms this notion, displaying a
non-zero offset in the EW scale (in \AA) for this line given by\\
\centerline{
\footnotesize{EW4541(ours) = ($-$0.096$\pm$0.06) + (1.01$\pm$0.11) $\cdot$
EW4541(M88).}}
Interestingly, no indication of any shift was found in the EW scale of
He I $\lambda$4471\\
\centerline{
\footnotesize{EW4471(ours) = ($-$0.01$\pm$0.06) + (1.01$\pm$0.1) $\cdot$
EW4471(M88).}}
These results imply that at least some of our targets might appear
as a somewhat earlier spectral type than denoted by \citet{mathys88}.
The data shown in Fig.~\ref{comp_EW} (lower panel) indicate
that our spectral types tend to be earlier, by half a subtype.
Though this discrepancy is comparable to the error budget of our
$quantitative$ classification, it might be significant because
of its
systematic character (see Section~\ref{conclusions}).
\begin{figure}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg6a.ps}}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg6b.ps}}
\caption{$Top$: Comparison of our high-resolution
equivalent widths (in \AA) of He~I~$\lambda$4471\, (asterisks) and He~II~$\lambda$4541
(diamonds) with those from \citet{mathys88}, for nine stars in common.
Dotted lines represent linear fits to the corresponding data. Note
the non-zero offset in the EW scale of He~II~$\lambda$4541\,.
$Bottom$: Comparison of the corresponding high-resolution
$quantitative$ spectral types. Spectroscopic binaries are highlighted
by either large circles (top panel) or filled diamonds (bottom panel).}
\label{comp_EW}
\end{figure}
Concerning luminosity classes, perfect agreement between the present
$quantitative$ luminosity classes and those attributed by \citet{mathys88}
was established for the nine stars in common.
The discrepancy between our present and the Mathys EW scale for
He~II~$\lambda$4541\, is difficult to interpret. On the one hand, binarity cannot
be an issue, because both double and single stars are involved.
On the other hand, a higher precision of our measurements can be
expected, because of the higher quality of our spectra. This higher
quality will allow us to consider the contribution of the broad Stark
wings more correctly. Given that in He~II~$\lambda$4541\,\, the wings are more
extended than in He~I~$\lambda$4471\,\, (linear against quadratic Stark effect),
we speculate that the effect of a higher $R$ and S/N on the
measured EWs will be more pronounced in the former than in the
latter line.
To conclude, the luminosity class criterion in the Conti
scheme is rather insensitive to changes in $R$, while the spectral
type indicator seems to be sensitive, though at a (very) low level.
\section{Correspondence between quantitative and morphological
spectral types} \label{quant-morph}
\begin{figure}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg7a.ps}}
\resizebox{\hsize}{!}
{\includegraphics[width=6cm,height=4cm]{15956fg7b.ps}}
\caption{$Top$: High-resolution $quantitative$ versus $morphological$
spectral types based on our high-resolution spectra degraded
to $R$=4\,000.
$Bottom$ High-resolution $quantitative$ versus high-resolution
$morphological$ spectral types.
}
\label{quant_old}
\end{figure}
It is important to verify the consisitency of the two alternative
schemes for the classification of O-type stars.
Earlier studies \citep{CL74, mathys88, mathys89} have shown
that $quantitative$ (either photographic or low-resolution
digital) and $morphological$ classification based on photographic
data are consistent to within {\it one to two subclasses}, with
the former being systematically later (on average, by roughly
half a spectral subtype).
The same comparison, but between the present $quantitative$
and the present $morphological$ classifications based on our degraded
spectra ($R$=4\,000), indicates a closer agreement (within $\pm$0.5
subtypes) without any systematic trend (Fig.~\ref{quant_old}, top panel).
When considering the correspondence between the high resolution
$quantitative$ and $morphological$ spectral types, we find
evidence in our data of a systematic difference, where
the former is now {\it earlier}, by up to 1.5 subtypes\footnote{This finding is
a natural consequence of our results from Sect.~\ref{results},
namely that within the high-resolution spectra He~I+II~$\lambda$4026\,/ He~I~$\lambda$4471\,\, can
appear deeper than He~II~$\lambda$4200\,/He~II~$\lambda$4541\,, although their EWs are smaller.}
(Figure~\ref{quant_old}, lower panel).
Thus, it appears that because of the higher quality of present-day
optical spectra (which increases the accuracy of quantitative
measurements and decreases the number of unresolved binary systems),
the consistency between the two classification schemes can be
significantly improved leading to practically equal results
(provided the $morphological$ classification is performed
using high-resolution spectra degraded down to $R$=4\,000).
\section{Summary and future perspectives.}
\label{conclusions}
Using high-resolution ESO/MTG FEROS spectra and applying the
Walborn and the Conti schemes, we have performed a spectral
classification of 19 Galactic O-type dwarfs and giants. We
have investigated the correspondence between these high-resolution
classifications and those attributed by Walborn \citep{walborn72,
walborn73} based on low-quality photographic spectra and those
assigned by \citet{mathys88} using low-resolution digital
spectra. In addition, we have also investigated the spectroscopic
status of our targets. The main results can be
summarised as follows.
\paragraph{1. \it Spectroscopic binaries.}
Using photometric/spectroscopic data from the literature in
combination with results from a {\sc fastwind} model atmosphere
analysis (Paper II), we confirm the binary nature of HD~152723
and HD~152247, provide strong evidence of
a companion in HD~93160 and HD~151003, and suspect possible
duplicity in HD~69464. Four of these stars are members of
cluster/associations and one is a field star.
Thus, the binary fraction among our sample is 26 percent if
both the cluster members and the field stars are considered.
Given that the majority of our targets has never been monitored
photometrically or spectroscopically, and that systems with
orbital periods too long for an easy detection and too short
for a direct angular separation can easily be lost, we
consider this estimate as a lower limit.
For spectral classification, our analysis
shows that binarity may or may not affect the derived
spectral type/luminosity class, depending on the magnitude
difference, separation, slit orientation etc. In any case,
and as might be expected, this effect does not lead to
systematic discrepancies (provided these systems are not used
as standards), but `only' increases the uncertainty
in the derived classification.
\paragraph{2. \it Correspondence between the present
$morphological$ classification and the Walborn results.}
Within a sample of 14 stars, we found evidence of
a systematic discrepancy between our $morphological$
spectral types and those attributed by Walborn, with the
former being later by up to 1.5 subtypes (Fig.~\ref{SPT_morph}).
This discrepancy is larger than the corresponding
uncertainties and thus significant. Concerning luminosity
classes, no evidence of any systematic difference between
our determinations and those from Walborn was found.
In our present understanding, the systematic discrepancy
in $morphological$ classification is due to an interplay
between stellar rotation and spectral resolution
(Figures~\ref{res_rot} and \ref{spectra})
as well as technical differences in the classification
process of photographic spectrograms and digital spectra
(Fig.~\ref{rbn}).
Another important outcome of our analysis
is that at least for narrow-lined stars ($v \sin i$\,$\le$100~\kms)
the ``rotational effect'' does not necessarily disappear when
the high-resolution spectra are degraded to the
resolution of the Walborn standards
(Figures~ \ref{res_rot} and \ref{resol_effect_2}).
\paragraph{3. \it Correspondence between the present
$quantitative$ classification and the Mathys results.}
Within a limited sample of nine stars,
we found that the spectral type criterion in the Conti
scheme, log~$W'$, is sensitive to changes in the spectral
resolving power (presumably due to effects on the EW of
He~II~$\lambda$4541\,), while the luminosity class criterion, log~$W''$,
is insensitive. In particular, the present high-resolution
$quantitative$ spectral types tend to be systematically earlier
(by typically half a subtype) than those attributed by Mathys.
Although comparable to the corresponding errors, the
established discrepancy might be significant because
it is systematic.
\paragraph{4. \it Correspondence between the present $quantitative$
and $morphological$ spectral types.} Although different
from earlier results, the modern-era $quantitative$
and $morphological$ classifications appear to be internally consistent
(provided that the $morphological$ classification is performed
using high-resolution spectra degraded to $R$=4\,000).
This result, if confirmed by better statistics, might help
us to
improve the accuracy and consistency of Galactic O star
classification after a corresponding new atlas of standard
stars at $R$=4\,000 has been set up and published (see below).
\smallskip
\noindent
From the results outlined above, one might conclude
that the $quantitative$ classification is more robust than the
$morphological$ one. This conclusion, however, would be
somewhat premature since it would be based on a very limited
sample of stars that furthermore does not include supergiants.
On the other hand, a spectral type discrepancy of even half
a subtype, if confirmed on the basis of higher quality statistics,
might be significant and therefore require
the Conti scheme to be updated using the power of high-quality
spectral observations and line profile simulations.
Such a thorough study is foreseen within our collaboration
in the ''{\it VLT-FLAMES
Tarantula Survey}" (PI: C.J. Evans).
Concerning the $morphological$ classification, a
possible solution of the problems identified in the
present study requires an update of the corresponding
approach. This revision might comprise the following
steps: (i) selection and observation (at relatively
high spectral resolution and S/N) of classification
standards, which will be used to create a new digital
atlas superseding the one of \citet{wf90};
(ii) development of a new classification technique that
allows the spectra of both unknown and standard
stars to be compared at the same resolution and $v \sin i$\,, and
(iii) reclassification of as many Galactic O-stars as
possible (observed uniformly and at the same $R$ as the
standards) to be used as a firm basis for future studies.
First steps in this direction have been undertaken
within the ``{\it Galactic O-stars Spectroscopic Survey}
(GOSSS)'' (PI: J. M. Appel\'aniz) and the ''{\it VLT-FLAMES Tarantula
Survey}''.
Within the GOSSS survey, a digital atlas of Galactic O-type
stars from both hemispheres will be created and published,
to supersede the one of \citet{wf90} in terms of both the
quality of the data (S/N $\approx$ 200, $R \approx$ 2\,500)
and the number
of standard stars included. Within the ``{\it VLT-FLAMES Tarantula
Survey}'' project, on the other hand, we plan to publish a
digital atlas of standard O-type stars using high-resolution ($R$
from about 40\,000 to about 80\,000), high signal-to-noise (S/N
$\approx$ 500) spectra collected from the ESO public archive and the
IACOB database \citep{simon10}. To diminish the combined effects of
resolution and rotation and ensure consistency with the modern-era
$quantitative$ classification, the spectra will be degraded
to
$R$=4\,000, and subsequently used to create an atlas of unprecedented
quality.
Finally, we point out that although updated with respect
to the resolution and rotational effects, the classification
of O-type stars might still be subject to significant third parameter
effects related to metallicity (see \citealt{markova09} and references
therein).
\acknowledgements{
We like to thank our anonymous referee for useful comments and
suggestions. N.M. gratefully acknowledges travel grants by the
IAC, Tenerife, Spain, and the Catania Observatory, Italy. This
investigation was supported by the Bulgarian NSF (contract DO 02-85).
SSD and AH acknowledge financial support from the Spanish Ministerio
de Ciencia e Innovaci\'on under the projects AYA2008-06166-C03-01
and the Consolider-Ingenio 2010. Program grant CSD2006-00070:
First Science with the GTC (http://www.iac.es/consolider-ingenio-gtc).
}
|
\section{Introduction\label{introduction}}
\input{introduction}
\section{The CCR and Second Quantization\label{ccr}}
\input{ccr}
\section{Prerequisites from Operator Theory\label{prerequisites}}
\input{prerequisites}
\section{The Number Operator\label{number}}
\input{number}
\section{General Quadratic Operators\label{operators}}
\input{operators}
\section*{Acknowledgment}
\input{acknowledgment}
\bibliographystyle{plain}
|
\section{Introduction}
Ultracold atomic gases are
proving to be extremely useful
tools for synthesizing low dimensional quantum models
owing to the huge degree of controlability they offer \cite{Bloch2008}.
The effective system dimensionality may be tuned in experiments
by manipulation of external trapping potentials,
with the underlying physics ultimately
set by the trap geometry and level of
quantum degeneracy \cite{Gorlitz2001}.
Increasing the trapping potential in one direction leads
to an effectively two-dimensional (2D) system
\cite{Gorlitz2001,Rychtarik2004,Stock2005,Smith2005},
which allows access to a number of interesting phenomena such as the
Berezinskii-Kosterlitz-Thouless transition and related studies on
the nature of the Bose gas in two-dimensions
\cite{Hadzibabic2006,Schweikhard2007,Kruger2007,Rath2008,Clade2009,Rath2010,Tung2010,Hung2011}.
Increasing the trapping potential in
a further direction
results instead in an effectively one-dimensional (1D) system
\cite{Moritz2003,Paredes2004,Kinoshita2004,Cacciapuoti2003,Trebbia2006,vanAmerongen2008,Armijo2010,Armijo2011,Esteve2006,
Dettmer2001,Hellweg2003,Gerbier2003,Richard2003,Manz2010,
Gustavson1997,Hinds2001,Schumm2005,Hofferberth2006,Fixler2007,Jo2007b,Hofferberth2007,Gross2010,Baumgartner2010,
Hinds1999,Ott2001,Hansel2001,Folman2002,Fortagh2007,Atomchips2011}.
In a 1D set-up,
one may obtain \cite{Petrov2000} either a weakly-interacting system, or,
for rather low densities,
a strongly-interacting Tonks-Girardeau gas \cite{Girardeau1960,Moritz2003,Paredes2004,Kinoshita2004}.
The finite temperature phase diagram of a {\em weakly interacting} 1D Bose
gas \cite{Petrov2000,AlKhawaja2003,Armijo2011} is more complex than that of a 3D
gas, due to a separation in the temperatures for the onset
of density and phase fluctuations.
Density fluctuations are typically
suppressed at higher temperatures than phase fluctuations, allowing
for the formation of a so-called quasi-condensate \cite{PopovBook}.
\begin{figure}[b!]
\centerline{
\psfrag{yaxis}{\scalebox{3.75}{$k_{B}T/\hbar\omega_{\perp}$}}
\psfrag{xaxis}{\scalebox{3.75}{$\mu/\hbar\omega_{\perp}$}}
\includegraphics[angle=0,scale=0.3,clip]{Cockburn_etal_fig1.eps}
}
\caption[]{
(Color online) Phase diagram indicating the theoretically obtained parameters
for the experiments considered. Hollow symbols indicate density profile
data of \cite{Trebbia2006} and \cite{vanAmerongen2008},
whereas filled symbols indicate the density fluctuation data of
\cite{Armijo2010}.
}
\label{fig:one}
\end{figure}
A number of experiments have recently probed the physics
of highly-elongated finite temperature Bose gases
at equilibrium,
including the direct observation and analysis of
density \cite{Cacciapuoti2003,Esteve2006,Armijo2010} and phase fluctuations
\cite{Dettmer2001,Hellweg2003,Gerbier2003,Richard2003,Manz2010}.
Understanding the
properties of matter waves in such geometries
is of key importance to
atom interferometers
\cite{Gustavson1997,Hinds2001,Schumm2005,Fixler2007,Jo2007b,Hofferberth2007,Gross2010,Baumgartner2010}
and atom chips \cite{Hinds1999,Hansel2001,Ott2001,Folman2002,Hofferberth2006,Fortagh2007,Atomchips2011}.
In this paper, we show that {\em in situ} density profiles and density fluctuations
from the elongated Bose gas experiments
of Trebbia \etal \cite{Trebbia2006},
van Amerongen \etal \cite{vanAmerongen2008}, and
Armijo \etal \cite{Armijo2010}, can be predicted {\em ab initio}
by means of an effective 1D stochastic model;
to achieve this, we propose and implement for the first time a modification
to the usual form of the stochastic Gross-Pitaevskii equation
\cite{Stoof1999,Stoof2001,Gardiner2003}
which additionally incorporates quasi-1D effects.
Such an extension beyond the purely 1D limit is required, since
these experiments probe regimes
for which both $\mu,\, k_{B} T \gtrsim \hbar \omega_{\perp}$,
as evident from Fig.\ref{fig:one}. Thus, one should in general
account for both a quasi-1D degenerate system exhibiting fluctuations
and for the non-negligible role of transverse thermal modes.
These are included here
by (i) implementing a stochastic quasi-1D equation of state for the
axial modes, and (ii) treating atoms in excited transverse modes
as independent, ideal Bose gases, as proposed in a related study
\cite{vanAmerongen2008}.
In Sec.~II we discuss our methodology in more detail, while Sec.~III focuses on
the {\it ab initio} reproduction of the experimental density profiles and
density fluctuation results reported in
\cite{Trebbia2006,vanAmerongen2008,Armijo2010},
with our conclusions presented in Sec.~IV.
\section{Methodology}
\label{sec:method}
We seek a model which can provide {\it ab initio} predictions for
experimentally measurable properties
obtained by {\it in situ} absorption imaging.
There are two issues that need to be addressed simultaneously here,
regarding the spatial extent of the quasi-condensate in the
transverse direction, and the role of atoms in excited transverse modes.
As mentioned, the important parameters affecting these are
the ratio of the chemical potential, $\mu$, and
thermal energy, $k_{B}T$, to the transverse ground state
energy $\hbar\omega_{\perp}$. For $\mu\ll\hbar\omega_{\perp}$,
the transverse ground state density has a Gaussian profile
of width $l_{\perp}$, the transverse oscillator length,
whereas for larger $\mu$,
this width becomes increased due to the
effect of repulsive interactions.
On the other hand, if $k_{B}T\ll\hbar\omega_{\perp}$,
then thermal occupation of the
transverse excited modes is negligibly small; for higher
temperatures, this is no longer true and atoms in these
modes will contribute considerably to experimental observables,
such as density profiles.
In the present work, we treat the
quasi-condensate and thermal modes separately:
consideration of transverse thermal modes is crucial for
matching total atom numbers and density profiles,
however they may be simply and accurately described as independent
equilibrium Bose gases, as shown in \cite{vanAmerongen2008};
fluctuating axial modes
are instead treated
within our modified stochastic model,
a novel feature of this work,
which is discussed in detail below.
{\em Quasi-condensate Modeling -}
The axial modes of a sufficiently elongated Bose gas are typically
subject to phase and density fluctuations due to thermal excitations with
wavelengths greater than the radial extent of the gas
\cite{Stringari1998,Petrov2001}. This
requires the inclusion of such fluctuations when describing
modes with energies
less than $\hbar\omega_{\perp}$. We therefore choose to describe these axial modes dynamically
using a
stochastic Gross-Pitaevskii equation (SGPE) \cite{Stoof1999,Stoof2001,Gardiner2003}.
In solving this equation, we make the following two assumptions:
(i) thermal modes (with energies
$>\hbar\omega_{\perp}$) may be treated as though at equilibrium,
and therefore represent a heat bath in contact with the axial sub-system
(these two components are assumed to be in diffusive and thermal
equilibrium with a temperature $T$ and chemical potential $\mu$);
(ii) the modes in the weakly trapped,
axial direction are sufficiently highly occupied that the classical
field approximation is valid
\cite{Svistunov1991,Stoof2001,Duine2001,Davis2001,Sinatra2001,Goral2001}.
Under assumptions (i) and (ii), the axial modes may be
represented by the 1D SGPE,
\begin{equation}
\begin{split}
i\hbar\frac{\partial \psi(z,t)}{\partial t}& = (1-i\gamma(z,t))
\bigg[-\frac{\hbar^{2}}{2m}\frac{\partial^{2}}{\partial z^{2}}+V(z)\\
&+g|\psi|^{2}-\mu\bigg]\psi(z,t)+\eta(z,t),
\end{split}
\label{eq:1dSGPE}
\end{equation}
where $\psi$ is a complex order parameter, $V(z)=m\omega_{z}^{2}z^{2}/2$
is the axial trapping potential, $g=2\hbar\omega_{\perp}a$
is the one-dimensional interaction strength (with $a$ the s-wave scattering length),
and $\eta$ is a complex Gaussian noise term, with correlations given by the relation
$\langle \eta^{*}(z,t) \eta(z',t') \rangle =2\hbar\gamma(z,t)k_{B}T\delta(z-z')\delta(t-t')$.
The strength of the noise, and damping, due to contact with the transverse thermal
modes, is given by $\gamma(z,t)$. This may be calculated {\it ab initio} in terms of the
Keldysh self-energy \cite{Stoof1999,Stoof2001,Duine2001}, however as we are interested only
in the system properties at equilibrium, we may approximate this quantity to be spatially and
temporally constant
\footnote{
Although not relevant for the equilibrium quantities probed here,
perturbations away from equilibrium would require a more
accurate calculation of $\gamma(z,t)$ including both spatial
\cite{Duine2001,Cockburn2010} and temporal dependence.
While a time-dependent thermal cloud has yet to be implemented within the SGPE
formalism, both temporal and spatial dependence of the damping parameter
are calculated fully self-consistently within a complementary `ZNG' scheme \cite{Zaremba1999},
whose validity is
however restricted to regimes of relatively small phase fluctuations.}.
To a good approximation, this is given by $\gamma = 3 \times 4ma^2 k_{B} T /(\pi\hbar^2)$
\cite{Penckwitt2002,Duine2004}.
Eq.\eqref{eq:1dSGPE} is valid in the scenario that the transverse
ground state is a Gaussian of width $l_{\perp}$. However,
as we wish
to consider experimental data from actual quasi-1D
systems, we should additionally modify
our model
in a manner which accounts for the transverse swelling of
the Bose gas due to repulsive interactions,
as seen experimentally \cite{Kruger2010}.
In the context of the ordinary Gross-Pitaevskii equation,
one may replace the 1D equation of state
$\mu[n]=g n$, where $n$ denotes the density,
with
\cite{Fuchs2003}
\begin{equation}
\mu[n]=\hbar\omega_{\perp}\left[\sqrt{1+4na}-1\right],
\label{eq:q1d_eos}
\end{equation}
which is obtained variationally by minimising the 3D GPE
with respect to the transverse chemical potential
\cite{Gerbier2004,Mateo2007,Frantzeskakis2010}.
This was shown in \cite{Gerbier2004} to interpolate smoothly across
the 1D-to-3D crossover \cite{Menotti2002},
and reduces to the 1D result
in the limit that $4an\ll1$,
as pointed out (for the ordinary GPE)
in \cite{Fuchs2003,Mateo2007,Frantzeskakis2010}.
Motivated by this we propose, somewhat heuristically,
a similar amendment to the 1D stochastic equation (Eq.\eqref{eq:1dSGPE});
for elongated but not truly 1D atomic clouds, this
gives rise to the modified stochastic equation
\begin{equation}
\begin{split}
i\hbar&\frac{\partial \psi(z,t)}{\partial t} = (1-i\gamma(z,t))
\bigg[-\frac{\hbar^{2}}{2m}\frac{\partial^{2} }{\partial z^{2}}+V(z)\\
&+\hbar\omega_{\perp}\left(\sqrt{1+4a|\psi|^{2}}-1\right)-\mu\bigg]\psi(z,t)+\eta(z,t)\;,
\end{split}
\label{eq:q1dSGPE}
\end{equation}
which we henceforth refer to as the quasi-1D SGPE.
The proposed modification to the 1D SGPE
is likely to become important when the inequality
$\mu\ll\hbar\omega_{\perp}$ is no longer satisfied,
signaling the onset of quasi-1D effects.
This method is of course only intended for modeling
very elongated systems with $\mu \lesssim{\rm few}~\hbar \omega_{\perp}$, in the
weakly-interacting
regime $mg/\hbar^2 n\ll1$; in this work we
thus restrict the application of this equation to the
weakly interacting regime, where numerous experiments exist.
In our stochastic scheme, the equilibrium state is reached in a dynamical manner,
when the effects of the noise
and damping terms of Eq.\eqref{eq:q1dSGPE} balance out.
Although
we demonstrate that this leads to
accurate equilibrium predictions,
the validity of this equation for
describing dynamical features remains to be investigated.
{\em Thermal Transverse Modes:}
Atoms in the transverse modes are considered to be in
static equilibrium, distributed according to
Bose-Einstein statistics; the relative importance of their contributions
depends on
the ratio $k_B T/\hbar\omega_{\perp}$,
and their contribution is significant in most experimentally relevant cases
\cite{Trebbia2006,vanAmerongen2008,Armijo2010}
(except at extremely low temperatures).
To account for their contribution to
total linear density profiles, we compute
(making the $\mu$ and $T$ dependence explicit for clarity)
\begin{equation}
n(z;\mu,T)=\langle|\psi(z;\mu,T)|^2\rangle+n_{\perp}(z;\mu,T)
\label{eq:ntot}
\end{equation}
where
\begin{equation}
n_{\perp}(z;\mu,T)=\frac{1}{\lambda_{\rm dB}}\sum_{j=1}^{\infty} (j+1)
{\rm g}_{1/2}\left[e^{(\mu-V(z)-j\hbar\omega_{\perp})/k_{B}T}\right],
\label{eq:nperp}
\end{equation}
${\rm g}_{1/2}[\ldots]$ is the polylogarithm (or Bose function)
of order $1/2$,
and $\lambda_{\rm dB}=h/\sqrt{2\pi m k_{B} T}$ is the
thermal de Broglie wavelength.
Such an addition of particles in thermal modes has already been used in
the so-called modified Yang-Yang model of
\cite{vanAmerongen2008}, where Eq.\ \eqref{eq:nperp} was used
in conjunction with a description of
atoms within axial modes based on Yang-Yang theory.
In our approach, we effectively replace the Yang-Yang theory with the SGPE,
thereby also providing an indirect comparison between SGPE and Yang-Yang,
in the weakly-interacting limit.
Furthermore, on making this replacement, the contribution given by $n_{\perp}$ remains
consistent with our treatment of bosons in excited transverse modes as having already
reached thermal equilibrium, as assumed within the SGPE model we apply here.
\begin{figure*}[ht]
\centering
\begin{tabular}{c}
\includegraphics[angle=0,scale=0.375,clip]{Cockburn_etal_fig2.eps}\\
\input{Cockburntab1.tab}
\end{tabular}
\caption[]{
(Color online) Total linear density profiles from the quasi-1D SGPE model
(Eq.\eqref{eq:ntot}; black solid line)
versus data from the experiment of Trebbia \etal \cite{Trebbia2006} (red circles).
Insets: quasi-1D SGPE contribution (Eq.\eqref{eq:q1dSGPE}; green dashed shaded region)
to the total linear density profiles (Eq.\eqref{eq:ntot}; black solid line).
Bottom: Table showing parameters $\mu$ and $T$ for the quasi-1D SGPE
density profiles which match the experimental data;
the parameters given in \cite{Trebbia2006} from a 3D Hartree-Fock
fit are also shown.
}
\label{fig:HFF}
\end{figure*
To summarize briefly, our approach for modeling
equilibrium properties of finite temperature
quasi-1D Bose gases is based on self-consistently solving Eq.\eqref{eq:q1dSGPE}
and Eq.\eqref{eq:nperp}
for the desired total atom number (or measured peak density) and temperature
\footnote{As the SGPE solutions can be somewhat time-consuming,
we find it convenient
(although by no means essential) to speed up the process by initially using the
modified Popov theory \cite{Andersen2002}
(which matches the SGPE results very well \cite{AlKhawaja2002,Cockburn2011}) to
arrive at a good initial condition for the self-consistent SGPE simulations.}.
In Sections \ref{sec:dprof} and \ref{sec:dfluct}, we
give a quantitative comparison between the proposed method
and published {\it in situ} experimental data,
thereby highlighting the usefulness of this method.
For a direct comparison to these experiments, we choose here
to fix the temperature to that reported
in the experimental papers, and then use $\mu$ as a free parameter, which we
vary until the required linear density is obtained at the trap centre.
\section{Comparison to Experiments}
\label{sec:comp}
\subsection{Density Profiles}
\label{sec:dprof}
We begin with a comparison of the model to total linear
density profiles, as obtained by {\it in situ} absorption
imaging within two experiments:
Sec. \ref{sec:dprof} gives a comparison to the data of Trebbia \etal \cite{Trebbia2006},
whose published analysis was based on a 3D Hartee-Fock model,
before discussing the measurements of van Amerongen \etal \cite{vanAmerongen2008},
who instead analyzed their results via the modified Yang-Yang theory.
In each study, experimental data was compared to theory in order to
quantify the equilibrium state through a
chemical potential and temperature.
The temperature may be quite straight-forwardly measured by fitting
the wings of the density distribution to the ideal gas result
(for sufficiently high temperatures), implying interactions
have little effect on measurements of this parameter.
Conversely, the chemical potential is far more dependent
on the model used in analyzing the
density profile; this is because different theories represent
the full quantum Hamiltonian of the interacting system
to different levels of approximation (see e.g. \cite{Proukakis2008}),
so incorporate many-body effects to a differing
degree.
It is worth pointing out also,
that the density profiles of the theory presented here,
and those from experimental
absorption imaging
share a common feature: namely, that
additional analysis is required to identify
a phase coherent (or `true' condensate)
and density coherent (or quasi-condensate)
fraction from the total density.
In the SGPE, the total density is due to both
coherent and incoherent particles,
however knowledge of the first and second order correlation functions
was shown to be sufficient to isolate both quasi-condensate and
true-condensate densities \cite{Cockburn2011,Wright2011}.
In particular, the method discussed in \cite{Cockburn2011}
may be easily applied to
directly extract such components from {\em experimental} measurements
of correlation functions,
thereby offering a more accurate, experimentally self-consistent,
characterisation of phase-fluctuating experiments (without the need to
resort to bimodal fits which become somewhat inaccurate in this limit).
\subsubsection{Comparison to work of Trebbia {\it et al.} \cite{Trebbia2006}}
\label{sec:Trebbia}
The results of Trebbia {\it et al.} demonstrated experimentally the breakdown
of the Hartree-Fock method when applied to highly elongated Bose gases \cite{Trebbia2006}.
It was concluded that this breakdown occurs because density fluctuations are not
accounted for accurately within this theory, since the energy lowering effect
of spatial correlations between atoms is not captured. These density correlations
\cite{Naraschewski1999,Prokofev2002,Proukakis2006b,Bisset2009b}
are key to
correctly predicting the onset of quasi-condensation and the associated
reduction in density
fluctuations; it was found, therefore, that the excited states did not saturate
within Hartree-Fock theory and so no quasi-condensate was predicted to form.
\begin{figure*}[ht]
\centering
\begin{tabular}{c}
\includegraphics[angle=0,scale=0.375,clip]{Cockburn_etal_fig3.eps}\\
\input{Cockburntab2.tab}
\end{tabular}
\caption[]{(Color online)
Top: Total linear density profiles from the quasi-1D SGPE (black solid line)
versus the experimental data of van Amerongen \etal \cite{vanAmerongen2008} (red circles).
Insets: Quasi-1D SGPE (thick black line) versus 1D SGPE (thin brown line)
density profiles.
Bottom: Table showing parameters $\mu$ and $T$ for the 1D and quasi-1D SGPE
density profiles which match the experimental data and modified Yang-Yang model
fits from \cite{vanAmerongen2008}.}
\label{fig:YY}
\end{figure*
If we compare instead to the results of the quasi-1D SGPE,
with the total linear density given by Eq.~\eqref{eq:ntot},
we see from Fig.\ref{fig:HFF} that the theoretical results (black solid line)
match well those obtained within the experiments (red circles).
The agreement is extremely good across the entire
range of temperatures considered, notably even at a temperature close to the
crossover from quasi-condensate to thermal gas (Fig. \ref{fig:HFF}(c)).
It is precisely this regime of critical fluctuations
in which a deviation from mean field theory might be expected, due to the
lack of a well defined mean field quantity.
Interestingly, behaviour suggestive of this was found
in \cite{Trebbia2006}
when comparing their data to the Hartree-Fock mean field model:
the experimental data was found to have a higher peak than
the mean field result [see Fig1(c) of \cite{Trebbia2006}],
which illustrates the potential importance of
including many-body effects,
as studied recently in the context of a
finite-temperature classical field theory \cite{Wright2011}.
We obtain a good
fit between the quasi-1D SGPE and experimental density profiles,
at the experimentally measured temperatures,
with a comparison between the chemical potentials extracted in our treatment
and the published values based on the
Hartree-Fock analysis of \cite{Trebbia2006} shown in the table
\footnote{Note that in comparing the SGPE data to
that of \cite{Trebbia2006}, we have:
(i) shifted the SGPE density profiles by $8\mu$m to the right
in order to match the position of the experimental peak densities;
(ii) used the relation $\mu=\mu_{\rm 3D}-\hbar\omega_{\perp}$, with
$\mu_{\rm 3D}$ the chemical potential reported in \cite{Trebbia2006}.}.
The deviation between the parameters of different theoretical methods
should not be of any concern, as it merely
highlights $\mu$ as a model dependent quantity,
i.e. it is dependent on the actual Hamiltonian used to analyse the experimental results,
and therefore
varies depending upon the level of
approximation \cite{Wright2011,Cockburn2011}.
The insets of Fig.\ref{fig:HFF} show the contributions to the total
linear density profiles due to the axial SGPE density (green dashed, shaded region),
with the remainder coming from Eq.\eqref{eq:nperp}.
The importance of the SGPE contribution is clear
in the first three plots, which shows an appreciable
fraction of atoms reside in axial modes for these parameters.
In the highest temperature case,
shown in Fig.\ref{fig:HFF}(d),
$\mu<0$ and the gas is entirely in the thermal phase;
here the density is instead due almost entirely to
atoms in tranverse modes.
\subsubsection{Comparison to work of van Amerongen {\it et al.} \cite{vanAmerongen2008}}
\label{sec:vanAmer}
The second experiment that we consider is that of van Amerongen {\it et al.}
\cite{vanAmerongen2008}.
This was the first experimental comparison to the exact Yang-Yang thermodynamic
solution to the finite temperature 1D Bose gas problem \cite{Yang1969},
also referred to as the thermodynamic Bethe ansatz.
In \cite{vanAmerongen2008}, the one-dimensional Yang-Yang theory was used to
represent the axial modes
and transverse ground state (of width $l_{\perp}$),
while the contribution to the linear density due
to atoms in transverse excited states, was accounted for using the method
we also adopt here. The total density profiles in \cite{vanAmerongen2008}
were therefore calculated using Eq.\eqref{eq:ntot},
with the role of the SGPE contribution $\langle|\psi|^{2}\rangle$
instead played by the 1D Yang-Yang prediction.
Therefore, in comparing to this work, we will gain insight on two fronts:
firstly, how well the SGPE matches the experimental data in this regime,
and simultaneously (but indirectly), how well the SGPE prediction for the density profiles
matches that due to Yang-Yang thermodynamics.
Fig.\ref{fig:YY} shows that the agreement between the
experimental data and the proposed quasi-1D SGPE approach is again very good across the entire
temperature range probed, including the crossover from quasi-condensate
to degenerate thermal gas.
We now wish to discuss how our quasi-1D SGPE results compare to those from the SGPE
with the usual 1D equation of state.
Practically, this means using the equilibrium result
of Eq.~\eqref{eq:1dSGPE}, rather than Eq.~\eqref{eq:q1dSGPE}, as the axial density
input $\langle|\psi|^{2}\rangle$ in Eq.~\eqref{eq:ntot}.
Although this approach also recovers closely the total density profiles
found with the quasi-1D SGPE at each temperature (insets to Fig. \ref{fig:YY}),
and therefore also those measured experimentally,
it is important to note that
each approach can
lead to slightly different chemical potentials
for the same temperature.
Importantly, we find that the parameters used
to obtain the 1D SGPE results are {\em identical} to those
obtained from fits of the modified Yang-Yang model to the density data in \cite{vanAmerongen2008}.
These results have also been reported
\cite{Kheruntsyan2010} to arise
within the context of the closely-related 1D stochastic
projected Gross-Pitaevskii equation (SPGPE)
\cite{Blakie2008} in parallel independent work,
which also looked at the momentum distribution of the gas
after focussing \cite{Shvarchuck2002}.
This provides an
indirect additional test between the 1D SGPE,
the 1D SPGPE
and Yang-Yang theories in the weakly-interacting regime.
The parameters predicted by the quasi-1D SGPE, 1D SGPE and modified Yang-Yang
models are shown in the Table of Fig.\ref{fig:YY}.
Having established that both the 1D and quasi-1D SGPE approaches
accurately reproduce experimental density profiles
(and therefore also those due to the modified Yang-Yang approach used in \cite{vanAmerongen2008}),
we now turn to an investigation of density fluctuations,
which provide a more sensitive probe for the validity of these theories.
\subsection{Density fluctuations}
\label{sec:dfluct}
Density fluctuations are increased markedly within an ideal Bose gas,
due to an effect of quantum statistics, which leads to
atomic bunching \cite{LandauLifshitz_statmech1}.
However, at sufficiently low temperatures, and in the presence of interactions,
quasi-condensation leads to atom-atom
correlations which overcome the tendency for bosonic atoms to bunch together,
and therefore to a reduction in the level of density fluctuations, relative to
those expected in an ideal Bose gas \cite{Esteve2006}.
\subsubsection*{Comparison to work of Armijo {\it et al.} \cite{Armijo2010}}
In a recent paper \cite{Armijo2010}, Armijo {\it et al.} measured the second and
third moments of the density fluctuations of a finite temperature Bose gas,
comparing these to theoretical predictions from ideal Bose gas
and quasi-condensate mean-field models,
and also the modified Yang-Yang model of \cite{vanAmerongen2008}.
We now briefly outline their experimental method
before describing the
numerical scheme we follow in order to
closely mimic this.
\emph{Experimental procedure:} A harmonic trapping potential
leads to a density profile in which the number of atoms varies spatially,
and therefore close to the trap centre it is possible to have a scenario
where there is sufficient levels of degeneracy such that a quasi-condensate is
formed, whereas in the low-density wings, the gas is still
effectively a non-interacting thermal gas.
Thus, at a single temperature, by scanning the spatial extent of the trapped gas,
it is possible to observe both the
enhancement of density fluctuations, due to
quantum statistics (low density, ideal Bose gas),
and their subsequent suppression, due to particle interactions
(higher density, quasi-condensate regime).
An approach based on this observation
was first undertaken experimentally in
\cite{Esteve2006}, and subsequently followed
by more detailed studies in \cite{Armijo2010,Armijo2011}.
In \cite{Armijo2010}, the gas was probed using a CCD camera, which
effectively divides observations of the gas into
pixel sized regions (of size $\Delta=4.5\mu m $ in this case).
Absorption imaging allowed for the number of atoms within each pixel,
$ N $, to be measured, and
repeated measurements provided a set of fluctuating values,
as well as an average
number per pixel, $\langle N \rangle$.
The $p$-th moment of the density fluctuations
for the set of measurements were then
calculated for each pixel as usual
via
$ \langle \delta N ^{p}\rangle=\left\langle \left( N - \langle N \rangle \right)^p \right\rangle$.
The aim of the present work is to demonstrate the SGPE as an {\it ab initio}
method for analyzing experimental findings,
and so we wish to implement a numerical scheme
which follows experimental procedures as closely as possible.
Fortunately, the grand canonical formulation of the SGPE
makes it relatively simple to simulate the experimental
methods used in \cite{Armijo2010}.
This is because, in addition to the
unified treatment of both (quasi-)condensate
and thermal atoms in density profiles, the SGPE shares a
second feature in common with experiments, namely
a shot-to-shot variation between individual realizations.
This variation enables us to straightforwardly model the equilibrium density
fluctuation experiments of \cite{Armijo2010},
but is also important in dynamical studies
\cite{Cockburn2010,Weiler2008,Damski2010,Das2011}.
\emph{Numerical SGPE procedure:}
Physical observables within the SGPE are obtained as products
of the stochastic wavefunction $\psi$, averaged over many realizations of
the noise $\eta$ (see Eq.~\ref{eq:q1dSGPE}, and Ref.~\cite{Cockburn2009} for a simple overview).
This leads naturally to a shot-to-shot variation
between numerical realizations,
in a way analogous to an individual experimental run.
By treating each noise realization like an experimental realization,
we are able to accurately synthesize the experimental procedure
of \cite{Armijo2010} within our numerical simulations.
We emphasise, however, that within this
analogy
it should be understood that single runs
represent weighted contributions to
averages over fluctuating quantities,
and that it is only such appropriately
averaged quantities that
we expect to match well those determined experimentally,
as indeed is found to be the case below.
Our numerical procedure is to
run a large number of stochastic simulations ($1000$),
each of which yields a fluctuating density profile.
We measure the number fluctuations within each pixel
sized region, by first spatially binning
the SGPE density data
into $\Delta$-sized regions \footnote{
The spatial numerical grid spacing, $\Delta z$, is chosen to give the
desired ultraviolet energy cutoff ($=\hbar\omega_{\perp}$),
which is much smaller
than the pixel size},
in order to give an output consistent
with that obtained in the experiments (see Fig.\ref{fig:method} (a)-(b)).
Integrating over the numerical grid points
within a single pixel yields $N_{z,\Delta}$,
the value for the (fluctuating) atom number from a single stochastic realization, within that pixel. Repeating the same procedure for
the mean total density generates a binned
average \emph{axial} pixel number, $\bar{N}_{z,\Delta}$.
The contribution from the axial modes to the second
($p=2$) and third ($p=3$) moments of the density fluctuations are
then calculated as
$ \langle \delta N ^{p} \rangle_{z}=\left\langle \left(N_{z,\Delta}-\bar{N}_{z,\Delta}\right)^p \right\rangle$.
Similarly, a binned transverse contribution to the average
pixel number $\bar{N}_{\perp,\Delta}$ may be obtained,
and the \emph{total} average atom number
in each pixel is then given by
$\bar{N}_{\Delta}=\bar{N}_{z,\Delta}+\bar{N}_{\perp,\Delta}$
(which corresponds to $\langle N \rangle$ in \cite{Armijo2010}).
\begin{figure}[bt!]
\centerline{
\includegraphics[angle=0,scale=0.255,clip]{Cockburn_etal_fig4.eps}
}
\caption[]
{
(Color online)
Top row: quasi-1D SGPE density (noisy red curve)
obtained from (a) a single numerical run (raw data),
and (b) corresponding spatially binned data; dashed black
curves display the density averaged over 1000 independent realisations.
(c) $\langle \delta N ^{2}\rangle_{m}$ \cite{endnote86} from the binned quasi-1D
SGPE data (black diamonds) and 1D SGPE data (green squares), compared against
correponding mean field results for an ideal Bose gas (brown solid line)
1D quasi-condensate (dotted maroon, horizontal) and quasi-1D quasi-condensate
(dashed red); the thin, vertical dashed lines indicate the `crossover' region
where the interaction and thermal energies become comparable. This data is for
$T=96$nK.
}
\label{fig:method}
\end{figure}
As we treat the atoms in the transverse modes in a static way, they give a
non-zero contribution only to average properties, and so do not
contribute to moments of the density fluctuations directly.
However, as we found in Section \ref{sec:dprof},
atoms in these modes are well approximated as a degenerate ideal gas, for which
$\langle \delta N ^{2}\rangle_{\perp} \simeq \langle \delta
N ^{3}\rangle_{\perp} \simeq \langle N \rangle_{\perp}$ \cite{Armijo2010},
and we therefore assume that these atoms contribute a factor
$\langle N \rangle_{\perp}$ to both second and third moments.
So, ultimately, we compute the total density fluctuations as
$\langle \delta N ^{p}\rangle=\langle \delta N ^{p}\rangle_{z}+\langle N \rangle_{\perp}$,
for $p=2,3$.
To illustrate our procedure, we plot in Fig.\ref{fig:method}
an example single run and average density
profile both before (Fig.\ref{fig:method}(a)), and after
(Fig.\ref{fig:method}(b)), spatial binning.
Comparing these plots, it is clear that the binning
procedure significantly smooths
the raw single run data, as would be
expected for this kind of a spatial
coarse graining procedure;
note that the binned data looks remarkably similar to the
experimental fluctuating density profile shown in \cite{Armijo2010}
(Fig. 1(c) of that work).
We additionally
show in Fig.\ref{fig:method}(c) the
variance in the density fluctuations which results
from
the set of $1000$
fluctuating binned densities; this is plotted
against the average number of atoms per pixel
\footnote{As detailed in \cite{Armijo2010}, we note there is a factor which
relates the experimentally measured moments,
$\langle\delta N\rangle_{m}$ to those obtained theoretically, via
$\langle\delta N ^{p}\rangle_{m} = \kappa_{p} \langle \delta N ^{p} \rangle$.
The factor $\kappa_{p}$ arises due to the finite spatial resolution of the
experiment, and therefore we must scale our findings
in order to account for this experimental issue.\label{endnote137}}.
Making use of the thermodynamic relation
$\langle\delta N ^2\rangle=k_{B}T\Delta(\partial n /\partial \mu)_T$
\cite{LandauLifshitz_statmech1},
it is possible to derive mean field results for the
density fluctuations,
based on both the ideal gas and quasi-condensate equations of state.
The ideal gas result is \cite{Armijo2011}
\begin{equation}
\langle\delta N ^2\rangle=\frac{1}{\lambda_{\mathrm{dB}}}
\sum_{j=1}^{\infty}\frac{\sqrt{j}}{k_{B}T}\frac{e^{j\mu/k_{B}T}}{\sqrt{j}}
\frac{1}{\left(1-e^{-j\hbar\omega_{\perp}/k_{B}T}\right)},
\end{equation}
while using the 1D equation of state
for a quasi-condensate ($\mu[n]=g n$)
gives the simple result \cite{Esteve2006}
$\langle\delta N ^2\rangle_{\rm 1D}={k_{B}T\Delta}/{g}$.
Using instead the quasi-1D equation of state, Eq.\eqref{eq:q1d_eos}, yields,
$\langle\delta N ^2\rangle_{\rm quasi-1D}=\langle\delta N ^2\rangle_{\rm 1D}\left[1+{(\mu-V(z))}/{\hbar\omega_{\perp}}\right]$.
Within the higher density region of Fig.\ref{fig:method}(c),
we show the mean field results due to the
1D and quasi-1D equations of state for a quasi-condensate.
It is clear that the 1D SGPE shows good agreement with the 1D mean field
result, whereas the quasi-1D SGPE instead agrees very well
with the quasi-1D mean field prediction.
The two vertical lines indicate the region
where
the interaction energy
and the average thermal energy become comparable.
For higher densities,
interactions significantly reduce
$\langle\delta N ^2\rangle$ below the ideal
gas prediction, and we see this is
more pronounced in the 1D SGPE case
relative to the quasi-1D SGPE data,
as expected from the 1D and quasi-1D mean field
predictions.
\begin{figure}[t!]
\centerline{
\includegraphics[angle=0,scale=0.29,clip]{Cockburn_etal_fig5.eps}
}
\caption[]
{
(Color online)
Second (top row: (a)-(b)) and third (bottom row: (c)-(d)) moments of the atom number
fluctuations from the quasi-1D SGPE data (black diamonds)
and experimental
data from the paper of Armijo \etal \cite{Armijo2010} (red circles).
Temperatures are $T=376$nK (left images) and $T=96$nK (right images).
Insets: axial (light blue crosses) versus axial
plus transverse (black diamonds) contributions to the
number fluctuations vs. the
ideal gas result (brown solid line).
}
\label{fig:Armijo}
\end{figure}
A key point from our analysis at this temperature,
is that while the ideal gas
equation of state is valid only for small densities,
and the mean-field quasi-1D equation of state
holds at high densities,
the quasi-1D SGPE, like the experimental data,
provides a smooth crossover between each of these regimes.
This is because, in moving outwards from the centre of the trap in the presence of
a quasi-condensate, at some point the gas changes phase to a thermal gas,
and a mean field theory cannot be expected to accurately describe
fluctuations in the transition region
near the edge of the quasi-condensate.
\emph{Comparison to Experiment:}
Fig.\ref{fig:Armijo} shows a comparison of the experimental
density fluctuation data obtained by Armijo {\it et al.} (red circles),
and the quasi-1D SGPE model (black diamonds).
We plot in the top row $\langle\delta N ^{2}\rangle_{m}$,
and in the bottom row $\langle\delta N ^{3}\rangle_{m}$,
each versus $\bar{N}_{\Delta}$ for two temperatures
($T=376$nK (left) and $96$nK (right)).
Concentrating on the $\langle\delta N ^{2}\rangle_{m}$ data first,
it is clear that the quasi-1D SGPE numerical results follow the
experimental data well in both the high temperature (left) and low
temperature cases (right images).
Notice that as the high-density, quasi-condensate regime is reached
in Fig.\ref{fig:Armijo}(b),
the 1D SGPE result (shown in Fig.\ref{fig:method}) would predict
too great a reduction in density
fluctuations relative
to the experimental results,
whereas the quasi-1D SGPE
captures the experimental behaviour very well.
Physically, this suggests that the effect of the transverse swelling
of the quasi-condensate near the centre of the trap cannot be ignored
for these parameters, and that the quasi-1D extension to the SGPE
is therefore essential here.
Armijo {\it et al.} found a similar trend when comparing between their
experimental data and the modified Yang-Yang model.
While the Yang-Yang result matched the low density data well,
at higher densities it displayed the same behaviour as
the 1D SGPE result.
Instead, Armijo {\it et al.} found the quasi-1D
mean field result to better reproduce
the experimental behaviour in the
higher density quasi-condensate regime.
So, as for the density profiles,
we again find the 1D SGPE data
to reproduce the 1D Yang-Yang behaviour, whereas the quasi-1D SGPE
goes beyond the modified Yang-Yang model by capturing the quasi-1D behaviour found in the experiment
in all regimes probed.
The results for the third moment (Fig.\ref{fig:Armijo}, bottom row),
$\langle\delta N ^{3}\rangle_{m}$, also show good agreement between the
quasi-1D SGPE and the experimental data (within the large experimental error bars).
The insets to Figs. \ref{fig:Armijo}(a)-(b) show clearly that the reduction
in density fluctuations compared to the ideal gas prediction (brown
solid line)
is more pronounced in the low temperature case, where there is a higher
quasi-condensate fraction. Contrary to this, the enhanced presence of thermal atoms
in the high temperature case,
provides an opportunity to test whether our approximate treatment of
the transverse mode contribution to $\langle\delta N ^{2}\rangle_{m}$,
leads to the correct behaviour at low density.
The results before (light blue crosses) and after (black diamonds)
the transverse contribution is added
are shown in the insets to Figs.\ref{fig:Armijo}(a)-(b).
This addition results in a noticeable shift upwards
in the data of Fig.\ref{fig:Armijo}(a)
while for the lower temperature
case, the difference following this addition is negligible.
Comparing the SGPE data to the expected low density result,
the ideal gas prediction for the same parameters
(solid brown line),
we see that adding a contribution $\langle N \rangle_{\perp}$ to
the axial contribution fully captures
the expected behaviour.
\section{Conclusions}
In conclusion, we have proposed and implemented
a suitably modified
one-dimensional stochastic Gross-Pitaevskii equation,
which was shown to provide excellent {\it ab initio} predictions
for both {\it in situ} experimental density profiles obtained by Trebbia \etal
\cite{Trebbia2006} and van Amerongen \etal \cite{vanAmerongen2008},
and {\it in situ} density fluctuation data from the experiment
of Armijo \etal \cite{Armijo2010}.
This was achieved by
matching peak densities (equivalent to total atom number)
to a hybrid scheme, which combines the aforementioned stochastic
model
with a previously reported approach based on
treating transverse thermal modes
as independent ideal Bose gases.
The study of density fluctuations showed that our combined approach
captures all experimental regimes studied in a unified manner, smoothly
interpolating between mean field models, whose individual validity is
restricted to either the low density or high density regimes.
Importantly, it was found that analyzing individual stochastic
realizations in the same way as individual experimental runs, led to
good agreement between the density statistics in each case.
Reducing our stochastic model to the previously tested one-dimensional stochastic
Gross-Pitaevskii equation showed that:
(i) the latter model is consistent with Yang-Yang predictions (in the weakly-interacting
regime probed here), and that
(ii) while both one-dimensional and quasi-one-dimensional approaches accurately
reproduce equilibrium density profiles, they do so with slightly different chemical
potentials.
The confidence gained from these successful comparisons suggests that the quasi-one-dimensional
stochastic Gross-Pitaevskii model, which was proposed here as a hybrid
of different previously implemented approaches, is an excellent model for describing quasi-condensate
experiments at equilibrium; the extent to which this model can be directly applied to study
dynamical features, such as experimentally-measured properties after expansion,
remains to be investigated.
\bigskip
\paragraph*{Acknowledgments.}
We would like to thank Isabelle Bouchoule, and Aaldert van Amerongen
and Klaasjan van Druten, for providing their experimental data for use
in our comparison.
NPP acknowledges discussions with Dimitri Frantzeskakis on the Gross-Pitaevskii equation
in the quasi-one-dimensional regime, and Matt Davis, Geoff Lee and Karen Kheruntsyan
for discussions on their related theoretical modeling.
We also thank Carsten Henkel and Antonio Negretti
for related discussions. This project was funded by the EPSRC.
\bibliographystyle{apsrev}
|
\section{Antimonotonic functions}
\label{antimonotonic}
\begin{definition}{\bf Antimonotonic boolean functions\\}
An antimonotonic boolean function on $2^{P_n}$ is a boolean-valued function $\alpha$ such that
\begin{equation}
\label{definition:antimonotonicfunction}
\forall S \subsetneq S' \subseteq P_n:\alpha(S') \Rightarrow \neg \alpha(S).
\end{equation}
We denote by $AMT(n)$ the set of antimonotonic boolean functions on $2^{P_n}$.
\end{definition}
\begin{theorem}
$\forall n > 0:|AMT(n)| = |MT(n)|$
\end{theorem}
\begin{proof}
Consider the boolean function $mt_n:AMT(n) \rightarrow MT(n)$ defined by
\[
\forall \alpha \in AMT(n):mt_n(\alpha) = \{S \subseteq P_n|\exists S' \in \alpha:S \subseteq S'\}
\]
The function $mt_n(\alpha)$ is monotonic by definition.
Given $f \in MT(n)$, $mt_n^{-1}(f)$ is given by the set of maximal elements of $f$.
\[
mt_n^{-1}(f) = \{S \in f|\neg \exists S' \in f:S \subset S'\}
\]
which is in $AMT(n)$ and is the inverse of $mt_n$ as can be verified.
\end{proof}
\noindent We will denote by $mt_n$ the mapping from $AMT(n)$ to $MT(n)$ and by $amt_n$ its inverse.
\\
\\
The ordering on $MT(n)$ induces an ordering on $AMT(n)$:
\begin{definition}{\bf Partial ordering of $AMT(n)$\\}
Given two antimonotonic boolean functions $\alpha, \beta \in AMT(n)$, $\alpha \le \beta$ iff $mt_n(\alpha) \le mt_n(\beta)$
and $\alpha < \beta$ iff $mt_n(\alpha) < mt_n(\beta)$.
\end{definition}
\begin{lemma}{\bf Partial ordering of $AMT(n)$\\}
The induced ordering on $AMT(n)$ is given by
\begin{equation}
\forall \alpha,\beta \in AMT(n):\alpha \le \beta \\
\Leftrightarrow \forall S \in \alpha:\exists S' \supseteq S:S' \in \beta
\label{eq:antimototonicordering}
\end{equation}
The strict ordering is given by
\begin{equation}
\forall \alpha,\beta \in AMT(n):\alpha < \beta \\
\Leftrightarrow \alpha \le \beta \wedge \exists S \in \beta:S \notin \alpha
\label{eq:strictantimototonicordering}
\end{equation}
\end{lemma}
\noindent The definition of immediate successors in $AMT(n)$ follows the corresponding definition for $MT(n)$.
As in $MT(n)$, $next(\alpha)$ is the set of immediate successors of $\alpha \in AMT(n)$.
An immediate successor is constructed according to the following lemma.
\begin{lemma}
\label{lemma:antimonotoneimmediatesuccessors}
Given $\alpha,\beta \in AMT(n)$, we have
\begin{equation}
\beta >_{im} \alpha \Leftrightarrow \alpha < \beta \ \wedge\ \exists S \in 2^{P_n}:\beta \backslash \alpha = \{S\} \wedge \forall X \subsetneq S:\exists X' \supseteq X: X' \in \alpha
\label{eq:antimonotoneimmediatesuccessionlemma}
\end{equation}
\end{lemma}
\begin{proof}
Let $\beta$ satisfy the condition of the lemma with $S$ the only set contained in $\beta$ and not in $\alpha$.
We prove that $mt_n(\beta) >_{im} mt_n(\alpha)$.
It is easily seen that $\alpha$ contains all elements of $\beta$ except $S$ and real subsets of $S$ with cardinality $|S|-1$.
So it follows that $mt_n(\beta) \backslash mt_n(\alpha) = \{S\}$.
Let $X \in mt_n(\beta)$ such that $X \not\subseteq S$.
$X$ clearly is in $mt_n(\alpha)$.
Let $X \not= S$ be a subset of S.
Again it follows that $X \in mt_n(\alpha)$.
So $mt_n(\beta)$ satisfies the conditions of lemma \ref{lemma:immediatesuccessors} and is an immediate successor
of $mt_n(\alpha)$.
It follows that $\beta >_{im} \alpha$.
This proves the leftward side of the equivalence. The proof of the other side is similar.
\end{proof}
\begin{definition}{\bf Extension of the notation\\}
Let $n_1,n_2 \in \mathbb{N},\ n_1\leq n_2$. We will use the notation $AMT(n_1,n_2)$ for the space of antimonotonic functions on subsets of $\{n_1,...,n_2\}$.
Clearly, we have $AMT(n) = AMT(1,n)$.
More generally, for any finite set $M$ of natural numbers, $AMT(M)$ will denote the the space of antimonotonic functions on subsets of $M$.
\end{definition}
\section{Conclusions}
\label{sec:conclusions}
We developed an interval algebra for the lattice of anti-monotonic functions on subsets of a finite set. Formulae for the intersection of intervals are given. An essential operator in the algebra is the external product operator of section \ref{sec:operators}.Together with the join operator, it allows for the decomposition as the union of disjoint intervals of the lattice of anti-monotonic functions on the subsets of a given finite set. This decomposition may be based on the elements of the sub-lattices defined on the sets accepted by any anti-monotonic function that covers the original set. A particularly interesting case of this decomposition relates to an anti-monotonic function in which all sets are disjoint. The external product operator then takes a particularly simple form. Given the rules for the intersection of intervals, the decomposition of the whole lattice naturally leads to a decomposition any interval. The partitions of the lattice as well as of its intervals resulting from this decomposition lead to a number of recursion formulae which can be used to count the number of elements in the lattice or interval (Dedekind problem). As an application of the algebra, we present an algorithm that enumerates all elements of the lattice in output-polynomial time. An implementation of this algorithm is available.
In appendix \ref{app:generallattices} we define the operator and prove the partitioning theorem for complete distributive lattices. Appendix \ref{app:example} describes the example of Young's lattice.
\section{Decomposition of the interval $\Upsilon_N$}
\label{sec:uniformspandecomposition}
The special intervals $[\alpha \vee \beta,\alpha \times \beta]$ for $\alpha,\beta \in AMT(N)$ from section \ref{sec:intervals} allow decomposing the interval $\Upsilon_N$, effectively reducing its dimensionality. That is why we refer to the following theorem as the ``General coordinate system'' theorem.
\begin{theorem}[General coordinate system]
\label{the:generalcoordinatesystem}
Given a finite set of positive integers $N$ and an antimonotonic function $\sigma$ with $sp(\sigma) = N$,
the set of nonempty intervals of the type
\[
{[\vee\{\kappa_S|S \in \sigma\},\times\{\kappa_S|S \in \sigma\}]}
\]
where $\{\kappa_S|S \in \sigma\}$ is any family of antimonotonic functions satisfying $\forall S \in \sigma:\kappa_S \in \Upsilon_S$, form a partition of $\Upsilon_N$.
\begin{proof}
In order to prove that the intervals from this set are disjoint, consider two families $\{\kappa_S|S \in \sigma\}$ and $\{\kappa'_S|S \in \sigma\}$. Let $S_1 \in \sigma$ be any set in the antimonotonic function $\sigma$ and let $\gamma \in {[\vee\{\kappa_S|S \in \sigma\},\times\{\kappa_S|S \in \sigma\}]} \cap {[\vee\{\kappa'_S|S \in \sigma\},\times\{\kappa'_S|S \in \sigma\}]}$ be any element in the intersection of the two intervals. We find
\begin{eqnarray}
\kappa_{S_1} & \le & \pi_{S_1}(\gamma) \le \kappa'_{S_1} \nonumber \\
\kappa'_{S_1} & \le & \pi_{S_1}(\gamma) \le \kappa_{S_1}
\end{eqnarray}
and conclude that $\kappa_{S_1} = \kappa'_{S_1}$ if the intervals have at least one element in common.
Two intervals are thus either equal or disjoint.
\\
\\
The union of all intervals is $\Upsilon_N$ since each $\gamma \in \Upsilon_N$ satisfies
\[
\gamma \in [\vee\{\pi_{S}(\gamma)|S \in \sigma\},\times \{\pi_{S}(\gamma)|S \in \sigma\}]
\]
Since for any $\gamma \in \Upsilon_N$, $sp(\gamma) = N$, we find that for any $S \in \sigma$, $sp(\pi_S(\gamma) )= S$ and hence $\pi_S(\gamma) \in \Upsilon_S$. This completes the proof.
\end{proof}
\end{theorem}
\begin{remark}
\label{rem:restrictioncoordinatesystem}
The restriction to antimonotonic functions $\sigma$ in Theorem \ref{the:generalcoordinatesystem} is not necessary for the theorem to hold.
If however, in the notation of Theorem \ref{the:generalcoordinatesystem}, $S_1 \subsetneq S_2$, we have
for $\kappa_{S_1} \in \Upsilon_{S_1}, \kappa_{S_2} \in \Upsilon_{S_2}$:
\begin{eqnarray}
\kappa_{S_1} \vee \kappa_{S_2} & \in & \Upsilon_{S_2} \nonumber \\
\kappa_{S_1} \times \kappa_{S_2} & = & \kappa_{S_1} \wedge \kappa_{S_2} \le \kappa_{S_1} \vee \kappa_{S_2} \nonumber
\end{eqnarray}
with equality only if $\kappa_{S_1} = \pi_{S_1}(\kappa_{S_2})$.
Consequently, any interval in the decomposition can be nonempty only if the latter condition is satisfied, leaving no freedom for the choice of $\kappa_{S_1}$.
\end{remark}
\begin{remark}
\label{rem:notnonemptycoordinatesystem}
The restriction does not remove all empty intervals from the decomposition.
If, still in the notation of Theorem \ref{the:generalcoordinatesystem}, two sets $S_1, S_2 \in \sigma$ are not disjoint and
\begin{equation}
\label{eq:interactioncoordinatesystem}
\pi_{S_1 \cap S_2}(\kappa_{S_1}) \neq \pi_{S_1 \cap S_2}(\kappa_{S_2})
\end{equation}
we find that
\[
\pi_{S_1 \cap S_2}(\kappa_{S_1} \vee \kappa_{S_2} ) \not\le
\pi_{S_1 \cap S_2}(\kappa_{S_1} \times \kappa_{S_2} )
\]
and consequently
\[
\kappa_{S_1} \vee \kappa_{S_2} \not\le
\kappa_{S_1} \times \kappa_{S_2}
\]
making an empty interval.
\end{remark}
\noindent Although the general form of Theorem \ref{the:generalcoordinatesystem} allows for studying symmetries in the decomposition,
the next theorem removes the empty intervals from Remark \ref{rem:notnonemptycoordinatesystem} by requiring $\sigma$ to contain only disjoint sets, i.e. to be a partition of $N$.
This restriction avoids the interaction expressed by condition (\ref{eq:interactioncoordinatesystem})
between the lower dimensional subspaces, hence its name ``orthogonal coordinate system theorem''.
\begin{theorem}[Orthogonal coordinate System]
\label{the:orthogonalcoordinatesystem}
For a finite set of positive integers $N$ and a partition $\sigma$ of this set, the set of intervals of the type
\[
{[\vee\{\kappa_S|S \in \sigma\},\times\{\kappa_S|S \in \sigma\}]}
\]
where $\{\kappa_S|S \in \sigma\}$ is any family of antimonotonic functions satisfying $\forall S \in \sigma:\kappa_S \in \Upsilon_S$,
form a partition of $\Upsilon_N$.
\begin{proof}
This theorem is a special case of Theorem \ref{the:generalcoordinatesystem}.
\end{proof}
\end{theorem}
\ \\
\noindent Given two finite sets $N1 \neq N2$, the corresponding intervals of uniform span $\Upsilon_{N_1}$ and
$\Upsilon_{N_2}$ are disjoint.
The family $(\Upsilon_S|S \subseteq N)$ forms a partition of $AMT(N) \backslash \{\emptyset\}$ and we have the disjoint union
\begin{equation}
AMT(N) = \{\emptyset\} \cup (\bigcup_{S \subseteq N} \Upsilon_S)
\label{eq:fulldecomposition}
\end{equation}
\begin{corollary}
The following expansion is an immediate consequence
\begin{equation}
\label{eq:uniformspanexpansion}
|AMT(n)| = 1
+ \binom{n}{0}|\Upsilon_{\emptyset}|
+ \binom{n}{1} |\Upsilon_{\{1\}}|
+ \binom{n}{2} |\Upsilon_{\{1,2\}}|
+ \dots
+ |\Upsilon_{\{1,\dots,n\}}|
\end{equation}
\end{corollary}
\noindent Equation (\ref{eq:fulldecomposition}) leads to the following decomposition of $AMT(N)$.
\begin{corollary}
\begin{equation}
AMT(N) = \{\emptyset\} \cup ( \bigcup_{(\kappa_S \in AMT(S)\backslash \{\emptyset\}|S \in \sigma)}[\vee \{\kappa_S|S \in \sigma\},\times \{\kappa_S|S \in \sigma \}])
\end{equation}
where the union is taken over all families of antimonotonic functions from $AMT(S)$, one for each $S \in \sigma$.
\end{corollary}
\begin{corollary}
A consequence of the decomposition in equation (\ref{eq:fulldecomposition}) is
\begin{equation}
|AMT(N)| = 1 + \sum_{S \subseteq N} |\Upsilon_S|
\end{equation}
\label{cor:fulldecomposition}
\end{corollary}
\noindent The two theorems allow to derive a number of recursion relations for Dedekind-like numbers.
The first one is an immediate consequence.
\begin{corollary}\label{cor:uniformspanintervalcounting}
For positive integers $n_1$ and $n$ such that $1 \le n_1 < n$ we have
\begin{equation}
|\Upsilon_{\{1,\dots, n\}}| = \sum_{\alpha \in \Upsilon_{\{1,\dots ,n_1\}},
\beta \in \Upsilon_{\{n_1+1,\dots, n\}}}{|[\alpha \vee \beta,\alpha \times \beta]|}
\end{equation}
\end{corollary}
\noindent Slight rearrangement in the sums allows to derive the following recursion relation.
\begin{corollary}\label{cor:intervalrecursioncounting}
For positive integers $n_1$ and $n$ such that $1 \le n_1 < n$ we have
\begin{equation}
|AMT(1,n)| = 1+\sum_{\alpha \in AMT(1,n_1)\backslash\{\emptyset\},
\beta \in AMT(n_1+1,n)\backslash\{\emptyset\}}{|[\alpha \vee \beta,\alpha \times \beta]|}
\end{equation}
\end{corollary}
\noindent An interesting special case of Corollary \ref{cor:intervalrecursioncounting} is $n_1 = n-1$:
\begin{corollary} For any positive integer $n > 1$
\begin{equation}
|AMT(1,n)| = \sum_{\alpha \in AMT(1,n-1)}{|[\emptyset,\alpha]|}
\label{cor:oneelementrecursioncounting}
\end{equation}
\begin{proof}
Let $n_1 = n - 1$. We have:
\begin{equation}
AMT(n_1+1,n) \backslash \{\emptyset\} = AMT(n,n) \backslash \{\emptyset\} = \{\{\emptyset\},\{n\}\}
\end{equation}
and Corollary \ref{cor:intervalrecursioncounting} implies
\begin{equation}
|AMT(1,n)| = 1+\sum_{\alpha \in AMT(1,n-1)\backslash\{\emptyset\},
\beta \in \{\{\emptyset\},\{\{n\}\}\}}{|[\alpha \vee \beta,\alpha \times \beta]|}
\end{equation}
Since $\{\emptyset\}$ is neutral for $\vee$ and $\times$, this is equivalent to
\begin{equation}
|AMT(1,n)| = |AMT(1,n-1)| + \sum_{\alpha \in AMT(1,n-1)\backslash\{\emptyset\}}{|[\alpha \vee \{\{n\}\},\alpha \times \{\{n\}\}]|}
\end{equation}
Each element $\kappa$ of $[\alpha \vee \{\{n\}\},\alpha \times \{\{n\}\}]$ can be written as $\alpha \vee (\kappa' \times \{\{n\}\})$ for some $\kappa' \in [\{\emptyset\},\alpha]$.\footnote{
$sp(\kappa) = sp(\alpha) \cup \{n\}$, $\kappa' = \pi_{sp(\alpha)}(\kappa \backslash \alpha)$.}
Hence
\begin{equation}
|[\alpha \vee \{\{n\}\},\alpha \times \{\{n\}\}]| = |[\{\emptyset\},\alpha]|
\label{cor:intervalsizeoneelement}
\end{equation}
and we find
\begin{eqnarray}
|AMT(1,n)| &=& |AMT(1,n-1)| + \sum_{\alpha \in AMT(1,n-1)\backslash\{\emptyset\}}{|[\{\emptyset\},\alpha]|} \nonumber \\
&=& 1 + \sum_{\alpha \in AMT(1,n-1)\backslash\{\emptyset\}}{|[\emptyset,\alpha]|} \nonumber \\
&=& \sum_{\alpha \in AMT(1,n-1)}{|[\emptyset,\alpha]|}
\end{eqnarray}
\end{proof}
\end{corollary}
\section{Structure of general intervals}
\label{sec:generaldecomposition}
For a finite set $N$ and $\alpha,\alpha',\beta,\beta' \in AMT(N)$, the intersection of intervals $[\alpha('),\beta(')]$ is given by
\begin{equation}
[\alpha,\beta] \cap [\alpha',\beta'] = [\alpha \vee \alpha',\beta \wedge \beta']
\end{equation}
The following theorem is an immediate consequence of Corollary \ref{cor:fulldecomposition}.
\begin{theorem}[Decomposition of intervals]
Given an interval $[\alpha,\omega]$ of antimonotone functions with $\alpha \neq \emptyset$ and an antimonotonic funtion $\sigma \le \omega$ with $sp(\sigma) = sp(\omega)$ then
the interval $[\alpha,\omega]$ is the disjoint union of intervals
\[
{[\vee\{\kappa_S|S \in \sigma\} \vee \alpha,\times\{\kappa_S|S \in \sigma\} \wedge \omega]}
\]
where each $\kappa_S \in [\pi_{S}(\alpha),\pi_{S}(\omega)]$.
\label{the:intervaldecomposition}
\end{theorem}
\section{Appendix}
\label{app:generallattices}
Let L be a complete distributive lattice with unit element 1.
Let $\alpha, \beta, \gamma \in L$ and define
\begin{itemize}
\item $Next(\alpha) = \{\kappa|\alpha <_{im} \kappa\}$
\item $base(\alpha) = \alpha \wedge(\vee(Next(1)))$
\item $top(\alpha) = \vee \{\kappa|base(\kappa) = base(\alpha)\}$
\item $\alpha \times \beta = \vee \{\kappa| top(\alpha) \wedge \kappa \le \alpha\ and\ \ top(\beta)\wedge\kappa \le \beta\}$
\end{itemize}
It is straightforward to prove the following
\begin{lemma}
\label{lem:externalproductandorder}
\[
\gamma \le \alpha \times \beta \Rightarrow \gamma \wedge top(\alpha) \le \alpha
\]
\begin{proof}
\begin{eqnarray*}
\gamma & \le & \vee \{\kappa| top(\alpha) \wedge \kappa \le \alpha \ and\ top(\beta)\wedge\kappa \le \beta\}\\
&\Leftrightarrow&\\
\gamma & = & \gamma \wedge (\vee \{\kappa| top(\alpha) \wedge \kappa \le \alpha\ and\ top(\beta)\wedge\kappa \le \beta\})\\
& = & \vee (\{\gamma \wedge \kappa| top(\alpha) \wedge \kappa \le \alpha\ and\ top(\beta)\wedge\kappa \le \beta\})\\
&\Rightarrow&\\
\gamma \wedge top(\alpha)& = & \vee (\{\gamma \wedge top(\alpha) \wedge \kappa| top(\alpha) \wedge \kappa \le \alpha\ and\ top(\beta)\wedge\kappa \le \beta\})\\
& \le & \alpha
\end{eqnarray*}
$\Box$
\end{proof}
\end{lemma}
\noindent Lemma \ref{lem:externalproductandorder} allows proving the following
\begin{theorem}
Let $\alpha, \beta \in L, \kappa_\alpha,\kappa'_\alpha \in [base(\alpha),top(\alpha)], \kappa_\beta,\kappa'_\beta \in [base(\beta),top(\beta)]$. We have
\begin{equation*}
[\kappa_\alpha \vee \kappa_\beta,\kappa_\alpha \times \kappa_\beta] \cap [\kappa'_\alpha \vee \kappa'_\beta,\kappa'_\alpha \times \kappa'_\beta] \neq \emptyset
\Leftrightarrow
\kappa_\alpha = \kappa'_\alpha,\kappa_\beta = \kappa'_\beta.
\end{equation*}
\begin{proof}
Let $\gamma \in [\kappa_\alpha \vee \kappa_\beta,\kappa_\alpha \times \kappa_\beta] \cap [\kappa'_\alpha \vee \kappa'_\beta,\kappa'_\alpha \times \kappa'_\beta] $. We find e.g.
\begin{eqnarray*}
\kappa_\alpha \le & \gamma \wedge top(\alpha) & \le \kappa'_\alpha\\
&and&\\
\kappa'_\alpha \le & \gamma \wedge top(\alpha) & \le \kappa_\alpha
\end{eqnarray*}
$\Box$
\end{proof}
\end{theorem}
\noindent Since for $\gamma \in [base(\alpha)\vee base(\beta), top(\alpha)\times top(\beta)],
\gamma \in [(\gamma \wedge top(\alpha)) \vee (\gamma \wedge top(\beta)), (\gamma \wedge top(\alpha)) \times (\gamma \wedge top(\beta))]$, we have constructed a partition of the interval
$[base(\alpha)\vee base(\beta), top(\alpha)\times top(\beta)]$.
\section{Appendix: an example}
\label{app:example}
Young's lattice is the lattice of Young diagrams. It is infinite. To apply our theory, let us initially use bounds for the horizontal $(n_h)$ and the vertical $(n_v)$ dimensions of the Young diagrams.
We denote by $vs_i$ vertical strip of size $i$ and by $hs_i$ the horizontal strip of size $i$.
The unit Young diagram is then $u = vs_1 = hs_1$.
The immediate successors of $u$ are $vs_2$ and $hs_2$.
We find that $vs_i \vee hs_j$ is an L-shaped Young diagram with vertical dimension $i$ and horizontal dimension $j$.
The external product $vs_i \times hs_j$ is a rectangular Young diagram with the same dimensions.
According to the decomposition theorem in appendix \ref{app:generallattices}, the set of all nonempty Young diagrams with the given bounds is given by the disjoint union
\begin{equation*}
\bigcup_{0<i<n_v,0<j<n_h}[vs_i \vee hs_j,vs_i \times hs_j].
\end{equation*}
Since this partition hods for all values of $n_v$ and $n_h$, the infinite lattice of nonempty Young diagrams is given by the disjoint union
\begin{equation*}
\bigcup_{0<i,0<j}[vs_i \vee hs_j,vs_i \times hs_j].
\end{equation*}
\section{Intervals of antimonotonic functions}
\label{sec:intervals}
For $\alpha \le \beta \in AMT(N)$ we define the intervals
\begin{equation}
{[\alpha,\beta]} = \{\kappa \in AMT(N)| \alpha \le \kappa \le \beta\}.
\end{equation}
\begin{prop}[Intersection of intervals]
For $\alpha,\beta,\alpha',\beta' \in AMT(n)$ the intersection of the intervals $[\alpha,\beta]$ and $[\alpha',\beta']$ is given by
\begin{equation}
[\alpha,\beta] \cap [\alpha',\beta'] = [\alpha \vee \alpha',\beta \wedge \beta']
\label{prop:intersectionofintervals}
\end{equation}
\end{prop}
\noindent For $\alpha,\beta \in AMT(n)$ we consider intervals of the form
\begin{equation}
[\alpha \vee \beta,\alpha \times \beta]
\end{equation}
Because $\emptyset$ is the annihilating element and $\{\emptyset\}$ the unit element for the external product
we have
\begin{equation}
[\alpha \vee \emptyset,\alpha \times \emptyset] = \emptyset,
[\alpha \vee \{\emptyset\},\alpha \times \{\emptyset\}] = \{\alpha\}
\label{prop:intervalemptyset}
\end{equation}
Since $sp(\alpha \vee \beta) = sp(\alpha \times \beta)$, we have
\begin{prop}[Uniform span]
For $\alpha, \beta \in AMT(n)$, each element $\kappa \in [\alpha \vee \beta,\alpha \times \beta]$ has the same span:
\begin{equation}
\forall \kappa \in [\alpha \vee \beta,\alpha \times \beta]:
sp(\kappa) = sp(\alpha \vee \beta)
\label{prop:uniformspan}
\end{equation}
\end{prop}
\noindent Since $\alpha = \pi_{sp(\alpha)}(\alpha) \leq \pi_{sp(\alpha)}(\alpha \vee \beta) $ and due to the inequality (\ref{prop:externalproductseparation}) for all elements in the interval $ [\alpha \vee \beta,\alpha \times \beta]$ the following invariant holds.
\begin{prop}
\label{prop:intervalinvariant}
For each $\kappa \in [\alpha \vee \beta,\alpha \times \beta]$ we have $\pi_{sp(\alpha)}(\kappa) = \alpha, \pi_{sp(\beta)}(\kappa) = \beta$.
\label{prop:intervalinequality}
\end{prop}
\noindent These definitions and properties are readily generalised for an arbitrary number of antimonotonic functions $\alpha_1,\dots,\alpha_k$ considering intervals of the type $[\alpha_1 \vee \dots \vee \alpha_k,\alpha_1\times \dots\times\alpha_k]$. To simplify the notation we will use
\begin{eqnarray}
\alpha_1 \vee \dots \vee \alpha_k &=& \vee\{\alpha_1,\dots,\alpha_k \} \nonumber \\
\alpha_1 \times \dots \vee \alpha_k &=& \times\{\alpha_1,\dots,\alpha_k \}
\end{eqnarray}
For a finite set $N$ of integers, we define
\begin{eqnarray}
{\alpha}_N &\equiv& \vee\{\{x\}|x \in N\} = \{\{x\}|x \in N\} \label{def:minalpha} \\
{\omega}_N &\equiv& \times\{\{x\}|x \in N\} = \{N\}\label{def:maxalpha}\\
{\Upsilon}_N & \equiv & [\alpha_N,\omega_N]
\end{eqnarray}
We find $sp({\alpha}_N) = sp({\omega}_N) = N$. The interval $\Upsilon_N = [{\alpha}_N,{\omega}_N]$ is the set of antimonotonic functions with span $N$. We investigate the structure of this interval in the next paragraph.
\section{Introduction}
The $n$th Dedekind number counts the number of antichains of subsets of an $n$-element set or the number of elements in a free distributive lattice on $n$ generators. Equivalently, it counts the number of monotonic functions on the subsets of a finite set of $n$ elements \cite{DEDEKIND,SLOANE}.
In 1969, Kleitman \cite{KLEITMAN} obtained an upper bound on the logarithm of the $n$th Dedekind number which was later improved by Kleitman and Markowsky \cite{KLEITMAN_MARKOWSKY} in 1975, namely
\[ (1+O((log n)/n)) \binom{n}{\lfloor n/2 \rfloor}.\]
In 1981, Korshunov \cite{KORSHUNOV} used a more complicated approach to give asymptotics for the $n$th Dedekind number itself. All these proofs were simplified by Kahn \cite{KAHN} in 2002 using an ``Entropy'' approach.
Finding a closed-form expression for the $n$th Dedekind number is a very hard problem, also known as Dedekind's problem and its exact values have been found only for $n \leq 8$ \cite{WIEDEMANN} :
\begin{eqnarray*}
2, 3, 6, 20, 168, 7581, &&7828354,\hfill\\
&& 2414682040998, 56130437228687557907788 .
\end{eqnarray*}
This is sequence A000372 in Sloane's Online Encyclopedia of Integer Sequences \cite{SLOANE}.
Monotonic boolean functions on sets of numbers are boolean-valued functions that preserve inclusion $(A \subset B \wedge f(B) \Rightarrow f(A))$.
A monotonic boolean function $f$ is uniquely determined by the largest subset $S$ for which $f(S) = true$.
These largest subsets define another category of functions which we will call antimonotonic.
These correspond to the antichains in the lattice defined by set inclusion \cite{COMTET}.
The main contribution of this paper is an algebra of intervals for the set of antimonotonic functions.
The algebra is based on the natural partial ordering on this set.
Given two comparable antimonotonic functions $\alpha \le \beta$ according to this ordering, the interval $[\alpha,\beta]$ is the set of antimonotonic functions $\gamma$ satisfying $\alpha \le \gamma \le \beta$.
We demonstrate how the space of antimonotonic functions on a finite set, and in fact any of its intervals, can be decomposed as a discrete union of intervals with border elements from lower dimensional subspaces.
This result is based on two decomposition theorems, the second being a consequence of the first.
The first theorem starts from a general antimonootonic function and uses the sets inside this function to generate
the lower dimensional subspaces.
The second theorem uses a partition of the basic set of n elements to genereate the antimonotonic function.
The latter theorem allows decompostion with somewhat different properties.
As applications of the decomposition, we derive a number of recursion formulas for Dedekind numbers.
These formulae are new to the best of our knowledge.
Another application is a class of algorithms for enumeration based on recursive sectioning of the intervals
We prove finiteness of these algorithms and argue that some of these algorithms are of output-polynomial time complexity.
In section \ref{ordering} we give a definition of the concept of monotonic functions, define partial order on monotonic function and characterise immediate succession.
In section \ref{antimonotonic} we define antimonotonic functions, give the well known isomorphism with the monotonic functions and derive the characterisation of the immediate succession for antimonotonic functions.
In section \ref{sec:operators} we define the well known join and meet operators as well as a convenient projection operator. A fourth operator is introduced which plays an important role in the decomposition theorems that follow.
Apart from the introduction of the latter operator, the first three sections mainly serve to set the notations.
In section \ref{sec:intervals}, we introduce our intervals of antimonotonic functions.
In section \ref{sec:uniformspandecomposition}, the decomposition of an important c lass of intervals - intervals of uniform span - is discussed.
In section \ref{sec:generaldecomposition}, the decomposition of general intervals is studied and an interval based enumeration algorithm is given.
We summarise our results in section \ref{sec:conclusions} and anticipate on further work.
\section{Partial ordering on monotonic functions}
\label{ordering}
Given a positive integer $n$, we denote by $P_n$ the set of the positive integers less than or equal to $n$.
We are interested in boolean functions defined on all subsets of $P_n$, which we denote as $2^{P_n}$.
Any such boolean function $f$ is uniquely defined by the set $f^{-1}(true)$.
We will not distinguish and consider $f$ as a function or as a set whichever is the clearest.
In other words, for a boolean function $f$ and a set $X \subseteq P_n$, we have
\begin{equation}
X \in f \text{ is equivalent to } f(X).
\label{eq:functionsassets}
\end{equation}
The following property defines monotonicity.
\begin{definition}{\bf Monotonic boolean functions\\}
A monotonic boolean function on $2^{P_n}$ is a boolean-valued function $f$ such that
\begin{equation}
\label{definition:monotonicfunction}
\forall S \subseteq S' \subseteq P_n:f(S') \Rightarrow f(S).
\end{equation}
We denote by $MT(n)$ the set of monotonic boolean functions on $P_n$.
\end{definition}
\noindent The set $MT(n)$ may be partially ordered by the following natural order relation.
\begin{definition}{\bf Partial ordering of $MT(n)$\\}
Given two monotonic boolean functions $f_1, f_2 \in MT(n)$, we define the partial order relation $\le$ as
\begin{equation}
f_1 \le f_2 \Leftrightarrow \forall S \subseteq P_n:f_1(S) \Rightarrow f_2(S)
\label{eq:mototonicordering}
\end{equation}
We define strict inequality in the usual way by
\begin{equation}
f_1 < f_2 \Leftrightarrow f_1 \not= f_2 \wedge f_1 \le f_2
\label{eq:strictmonotonicordering}
\end{equation}
We will also use $\ge (>)$:
\begin{equation}
f \ge (>) g \Leftrightarrow g \le (<) f
\label{eq:partialgreaterthan}
\end{equation}
\end{definition}
\noindent The immediate successors of a function in $MT(n)$ are defined by
\begin{definition}{\bf Immediate succession in $MT(n)$\\}
Given functions $f, g \in MT(n)$, we say that $g$ is an immediate successor of $f$ iff
\begin{equation}
f < g \wedge \nexists h \in MT(n):f < h < g
\label{eq:immediatesuccession}
\end{equation}
We denote the set of immediate successors of $f \in MT(n)$ by $Next(f)$ and if $g \in Next(f)$ we say $f <_{im} g$ or equivalently $g >_{im} f$.
\end{definition}
\noindent The following lemma allows to construct immediate successors in $MT(n)$.
\begin{lemma}
\label{lemma:immediatesuccessors}
Given $f,g \in MT(n)$, we have
\begin{equation}
f <_{im} g \Leftrightarrow \exists S \in 2^{P_n}:g \backslash f = \{S\} \wedge \forall X \subsetneq S:X \in f
\label{eq:immediatesuccessionlemma}
\end{equation}
\end{lemma}
\begin{proof}
Clearly $|g \backslash f| > 1 \Rightarrow g \notin Next(f)$ and if $g \backslash f = \{S\}$ with $S' \notin f$ for some $S' \subsetneq S \subset P_n$, then $f < f \cup 2^{S'} < g$ and $g \notin Next(f)$. This proves the ``$\Rightarrow$'' part of the lemma.
Given $f \in MT(n)$, $S \notin f$ such that $\forall X \subsetneq S:X \in f$.
$f \cup \{S\}$ is monotonic and thus in $MT(n)$. We prove that $(f \cup \{S\}) >_{im} f$.
Suppose there is a $h \in MT(n)$ with $f < h < (f \cup \{S\})$.
$h < (f \cup \{S\})$ implies that $\forall X \in h: X \in f \vee X = S$.
Since $f < h$, $X \in f$ implies $X \in h$ so that $(h \not= f \cup \{S\} \Rightarrow S \notin h)$.
But this would imply that $f = h$, a contradiction.
\end{proof}
\noindent Since the smallest monotonic function is $\emptyset$ and the largest is $2^{P_n}$, we have
\begin{corollary}
The length of the longest chains of immediate successors is $2^n+1$.
\end{corollary}
\subsection{Ranks and distances}
One of the main results in this article is a recursive procedure to split an interval in ever smaller subintervals to arrive at an enumeration of its elements.
We need one more device before we can present the procedure.
The device is a measure for the distance between two antimonotonic functions.
We will use this measure for a heuristic estimate of the size of an interval,
and to ensure that after the split, the fragments are smaller than the whole.
We first introduce the rank of an antimonotonic function. It is defined as the number of different subsets of sets in the function:
\begin{definition}
\begin{equation}
\forall \alpha \in AMT(N):
rank(\alpha) = \sum_{A \in \alpha}{2^{|A|}}
- \sum_{A\neq B \in \alpha}{2^{|A \cap B|}}
+ \sum_{A\neq B\neq C \in \alpha}{2^{|A \cap B \cap C|}} -...
\label{def:rankfunction}
\end{equation}
\end{definition}
\noindent The distance between two antimonotonic functions is defined as
\begin{definition}
\begin{equation}
\forall \alpha,\beta \in AMT(N):
d(\alpha,\beta) = rank(\alpha) + rank(\beta) - 2\times rank(\alpha\wedge\beta)
\label{def:distancefunction}
\end{equation}
\end{definition}
\subsection{Recusively partitioning intervals}
The decomposition in Theorem \ref{the:intervaldecomposition} can be repeated recusively to enumerate antimonotonic functions over a complete space or within an interval.
It is not hard to see that for $\emptyset < \alpha < \omega$, there is always a set $N \subseteq sp(\omega)$
allowing to produce smaller intervals.
\begin{prop}
For antimonotonic functions $\emptyset < \alpha < \omega$,
there is always a nonempty set $N \subseteq sp(\omega)$ such that
$\pi_{N}(\alpha) < \pi_{N}(\omega)$. If $|\omega| > 1$ or $\alpha$ is not an immediate predecessor of $\omega$,
there is such a set satisfying $N \subsetneq sp(\omega)$
\label{prop:descentproperty}
\begin{proof}
$N=sp(\omega)$ satisfies the first condition. Another example is found as follows. Let $A \in \omega$ be a non-empty set such that $A \notin \alpha$.
It can be seen that $\pi_A(\alpha) < \pi_A(\omega)$.
(Should $P_A(\alpha) = P_A(\omega)$, then there would be a set $B \in \alpha$
with $A \subsetneq B$ contradicting $\alpha < \omega$.)
If $|\omega| > 1$ then $|sp(\omega)| > |A|$ and $sp(\omega) \backslash A \not= \emptyset$.
If $|\omega| = 1$ and $\alpha$ is not an immediate predecessor of $\omega$
then there must be a set $X \notin \alpha$ of the form $sp(\omega) \backslash \{a\}$ with $a$ an element of the only set in $\omega$.
We find that $P_X(\alpha) < P_X(\omega)$ and $sp(\omega) \backslash X$ is non empty.
\end{proof}
\end{prop}
\noindent Given that $\alpha <_{im} \omega \Leftrightarrow |[\alpha,\omega]| = 2$,
and building on Property \ref{prop:descentproperty},
Procedure \ref{proc:intervallisting} lists the elements of an interval $[\alpha,\beta]$.
\begin{algorithm}
\caption{List all elements of an interval of antimonotonic functions and return their number}
\begin{algorithmic}
\label{proc:intervallisting}
\REQUIRE $\alpha \not= \emptyset, \beta \not= \emptyset$
\ENSURE $|[\alpha,\beta]|$
\ENSURE print all elements of $[\alpha,\beta]$
\STATE {\bf function} listElements($\alpha,\beta \in AMT(n)$) {\bf returns} number
\IF{$\alpha \not\le \beta $}
\RETURN 0
\ENDIF
\IF {$\alpha = \beta$ }
\PRINT $\alpha$
\RETURN 1
\ENDIF
\IF {$|b| = 1\ and\ \alpha <_{im} b$}
\PRINT $\alpha, \beta$
\RETURN 2
\ENDIF
\STATE select disjoint subsets X,Y of $sp(\beta)$ such that
\STATE $d(P_X(\alpha),P_X( \beta))\ge 1\ and\ d(P_Y( \alpha),P_Y( \beta)) \ge 1$
\STATE {\bf set} COUNT = 0
\FORALL{$\kappa \in [P_X(\alpha),P_X(\beta)]$}
\FORALL{$\lambda \in [P_Y(\alpha),P_Y(\beta)]$}
\STATE {\bf set} COUNT = COUNT + listElements($(\kappa \vee \lambda) \vee \alpha, (\kappa \times \lambda) \wedge \beta $)
\ENDFOR
\ENDFOR
\RETURN COUNT
\STATE {\bf end function}
\end{algorithmic}
\end{algorithm}
Note that the recursion is not only over the fragment $[(\kappa \vee \lambda) \vee \alpha,(\kappa \times \lambda)\wedge \beta]$ but also in the iterations over $[P_X(\alpha),P_X(\beta)]$ and $[P_Y(\alpha),P_Y(\beta)]$.
Since each couple $\kappa, \lambda$ selected from these intervals are used in the construction of the elements of the new interval, and since finding a suitable split is linear in the size of $span(\omega)$, this procedure is of complexity $|span(\omega)|*outputsize$. Given the double exponential size of the output with respect to $span(\omega)$, we can say that the procedure is essentially linear in the size of the output.
\noindent The conditions $d(P_X(\alpha),P_X( \beta)) \ge 1$ and $d(P_Y( \alpha),P_Y( \beta)) \ge 1$ can be replaced by a condition that splits the interval in parts that are as equal as possible. This reduces the depth of the recursion and has a beneficial effect on the construction of the intervals (bigger $\kappa, \lambda$). It does bring a cost however in the computation of the split of $span(\omega)$. It is not immediately clear what the impact on the complexity is.
\section{Projection and other operators}
\label{sec:operators}
\subsection{Projection}
Given two finite sets $N' \subseteq N$ the projection from the space $AMT(N)$ to $AMT(N')$ is defined by
\begin{equation}
\pi_{N'}:AMT(N) \rightarrow AMT(N'):\alpha \rightarrow sup(\{A \cap N'|A \in \alpha\})
\label{def:projection}
\end{equation}
\begin{prop}[Order conservation by projection]
For sets of integers $N' \subseteq N$, $\alpha, \beta \in AMT(N)$ we have
\begin{equation}
\alpha \le \beta \Rightarrow \pi_{N'}(\alpha) \le \pi_{N'}(\beta)
\label{prop:orderconservationprojection}
\end{equation}
\end{prop}
\noindent Three further operators are useful to explore the space of antimonotonic functions further.
The first two immediately follow from the meet and join operators in the lattice of anti-chains of which the anti-monotonic functions are a representation.
The definition of the third operator given here is tied to the specifics of anti-monotonic functions, but as we demonstrate in appendix \ref{app:generallattices}, this operator and its applications in the intervals we subsequently derive can be readily generalised to complete distributive lattices.
\subsection{Meet}
We define the idempotent meet operator $(\wedge)$ as follows
\begin{equation}
\forall \alpha,\beta \in AMT(N):\alpha \wedge \beta = sup(\{A \cap B| A \in \alpha, B \in \beta\}))
\label{def:externaldot}
\end{equation}
Where de operation $sup$ is defined as follows
\begin{equation}
\forall S \subset 2^N:sup(S) = \{A \in S|\nexists A' \in S: A \subsetneq A'\}
\label{def:sup}
\end{equation}
Obviously we have that $\forall \alpha,\beta \in AMT(N):\alpha \wedge \beta \in AMT(N)$.
The following property makes the connection with the more general operator for anti-chains.
\begin{prop}[Largest common lower bound]
For antimonotonic functions $\alpha,\beta,\kappa$ we have
\begin{equation}
\kappa \le \alpha\ and\ \kappa \le \beta \Leftrightarrow \kappa \le \alpha \wedge \beta
\label{prop:commonlowerbound}
\end{equation}
\begin{proof}
($\Rightarrow$) For any $K \in \kappa$ we have sets $A \in \alpha, B \in \beta$, such that $K \subseteq A \cap B$.\\
($\Leftarrow$) For any $K \in \kappa$, pick $A \cap B \in \alpha \wedge \beta$ such that $K \subseteq A \cap B$. Clearly $K \subseteq A$ and $K \subseteq B$.
\end{proof}
\end{prop}
$\alpha \wedge \beta$ is the largest anti-monotonic function that is smaller than both $\alpha$ and $\beta$.
\subsection{Join}
We define the idempotent join operator $\vee$ as follows
\begin{equation}
\forall \alpha,\beta \in AMT(1,n):\alpha \vee \beta = sup(\alpha \cup \beta)
\label{def:externalsum}
\end{equation}
$\alpha \vee \beta$ is antimonotonic.
One finds for $\alpha \neq \emptyset$:
\begin{eqnarray}
\alpha \vee \emptyset & = & \alpha \nonumber\\
\alpha \vee \{\emptyset\} & = & \alpha
\label{prop:externalsum}
\end{eqnarray}
\begin{prop}[Least common upper bound]
For antimonotonic functions $\alpha,\beta,\kappa$ we have
\begin{equation}
\kappa \ge \alpha\ and\ \kappa \ge \beta \Leftrightarrow \kappa \ge \alpha \vee \beta
\label{prop:commonupperbound}
\end{equation}
\end{prop}
$\alpha \vee \beta$ is the smallest anti-monotonic function dominating both $\alpha$ and $\beta$.
\subsection{External product}
We define the span of an anti-monotonic function $\alpha$ as
\begin{definition}
$sp(\alpha) = \cup_{A \in \alpha}{(A)}$
\end{definition}
The external product with respect to subsets of $N$ is defined as follows
\label{sec:externalproduct}
\begin{definition}
Let $N$ be a set of integers,
$\alpha,\beta \in AMT(N)$
\noindent The external product of $\alpha$ and $\beta$ is given by
\begin{equation}
\alpha \times \beta = max\{\kappa \in AMT(N): sp(\kappa) = sp(\alpha) \cup sp(\beta), \pi_{sp(\alpha)}(\kappa) \le \alpha, \pi_{sp(\beta)}(\kappa) \le \beta\}
\label{def:externalproduct}
\end{equation}
\end{definition}
\noindent The maximum in definition (\ref{def:externalproduct}) is unique as is shown by the construction
\begin{equation}
\alpha \times \beta = sup\{(A \backslash sp(\beta)) \cup (B \backslash sp(\alpha)) \cup (A \cap B)|A \in \alpha, B \in \beta\}
\label{def:externalproductb}
\end{equation}
The operation $\times$ has the associative and commutative properties. It is idempotent. Its neutral element is $\{\emptyset\}$ and $\emptyset$ is the annihilating element.
The following important property is an immediate consequence of the definition.
\begin{prop}
For $\alpha,\beta,\gamma \in AMT(N)$ we have
\begin{equation}
\label{prop:externalproductseparation}
\gamma \le \alpha \times \beta \Rightarrow \pi_{sp(\alpha)}(\gamma) \le \alpha
\end{equation}
\end{prop}
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.